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 246 毫秒
1.
We performed sensory evaluations on 141 bottles of sake and analyzed the relationship between the d-amino acid concentrations, and the taste of the sake using principal component analysis, which yielded seven principal components (PC1–7) that explained 100 % of the total variance in the data. PC1, which explains 33.6 % of the total variance, correlates most positively with strong taste and most negatively with balanced tastes. PC2, which explains 54.4 % of the total variance, correlates most positively with a sweet taste and most negatively with bitter and sour tastes. Sakes brewed with “Kimoto yeast starter” and “Yamahaimoto” had high scores for PC1 and PC2, and had strong taste in comparison with sakes brewed with “Sokujo-moto”. When present at concentrations below 50 μM, d-Ala did not affect the PC1 score, but all the sakes showed a high PC1 score, when the d-Ala was above 100 μM. Similar observations were found for the d-Asp and d-Glu concentrations with regard to PC1, and the threshold concentrations of d-Asp and d-Glu that affected the taste were 33.8 and 33.3 μM, respectively. Certain bacteria present in sake, especially lactic acid bacteria, produce d-Ala, d-Asp and d-Glu during storage, and these d-amino acids increased the PC1 score and produced a strong taste (Nojun). When d- and l-Ala were added to the sakes, the value for the umami taste in the sensory evaluation increased, with the effect of d-Ala being much stronger than that of l-Ala. The addition of 50–5,000 μM dl-Ala did not effect on the aroma of the sakes at all.  相似文献   

2.
A recombinant l-fucose isomerase from Caldicellulosiruptor saccharolyticus was purified as a single 68 kDa band with an activity of 76 U mg?1. The molecular mass of the native enzyme was 204 kDa as a trimer. The maximum activity for l-fucose isomerization was at pH 7 and 75°C in the presence of 1 mM Mn2+. Its half-life at 70°C was 6.1 h. For aldose substrates, the enzyme displayed activity in decreasing order for l-fucose, with a k cat of 11,910 min?1 and a K m of 140 mM, d-arabinose, d-altrose, and l-galactose. These aldoses were converted to the ketoses l-fuculose, d-ribulose, d-psicose, and l-tagatose, respectively, with 24, 24, 85, 55% conversion yields after 3 h.  相似文献   

3.
Astrocytic excitatory amino acid transporters (EAATs) regulate excitatory transmission and limit excitotoxicity. Evidence for a functional interface between EAATs and glial fibrillary acidic protein (GFAP) relevant to astrocytic morphology led to investigations of actions of transportable (d-Aspartate (d-Asp) and (2S,3S,4R)-2-(carboxycyclopropyl)glycine (l-CCG-III)) and non-transportable (dl-threo-β-benzyloxyaspartate (dl-TBOA)) inhibitors of Glu uptake in murine astrocytes. d-Asp (1 mM), l-CCG-III (0.5 mM) and dl-TBOA (0.5 mM) produced time-dependent (24–72 h) reductions in 3[H]d-Asp uptake (approximately 30–70%) with little or no gliotoxicity. All drugs induced a profound change in phenotype from cobblestone to stellate morphology and image analysis revealed increases in the intensity of GFAP immunolabelling for l-CCG-III and dl-TBOA. Cytochemistry indicated localized changes in F-actin distribution. Cell surface expression of EAAT2, but not EAAT1, was elevated at 72 h. Blockade of Glu uptake by both types of EAAT inhibitor exerts longer-term effects on astrocytic morphology and a compensatory homeostatic rise in EAAT2 abundance.  相似文献   

4.
The d,d-transpeptidase activity of Penicillin Binding Proteins (PBPs) is essential to maintain cell wall integrity. PBPs catalyze the final step of the peptidoglycan synthesis by forming 4 → 3 cross-links between two peptide stems. Recently, a novel β-lactam resistance mechanism involving l,d-transpeptidases has been identified in Enterococcus faecium and Mycobacterium tuberculosis. In this resistance pathway, the classical 4 → 3 cross-links are replaced by 3 → 3 cross-links, whose formation are catalyzed by the l,d-transpeptidases. To date, only one class of the entire β-lactam family, the carbapenems, is able to inhibit the l,d-transpeptidase activity. Nevertheless, the specificity of this inactivation is still not understood. Hence, the study of this new transpeptidase family is of considerable interest in order to understand the mechanism of the l,d-transpeptidases inhibition by carbapenems. In this context, we present herein the backbone and side-chain 1H, 15N and 13C NMR assignment of the l,d-transpeptidase from Bacillus subtilis (LdtBs) in the apo and in the acylated form with a carbapenem, the imipenem.  相似文献   

5.
A novel protodioscin-(steroidal saponin)-glycoside hydrolase, named protodioscin-glycosidase-1 (PGase-1), was purified and characterized from the Aspergillus oryzae strain. The molecular mass of this enzyme was determined to be about 55 kDa based on SDS-polyacrylamide gel electrophoresis. PGase-1 was able to hydrolyze the terminal 26-O-β-d-glucopyranoside of protodioscin (furostanoside) to produce dioscin (spirostanoside), and then further hydrolyze the terminal 3-O-(1?→?4)-α-l-rhamnopyranoside of dioscin to form progenin III. However, PGase-1 could hardly hydrolyze the 3-O-(1?→?2)-α-l-rhamnopyranoside of progenin III, 3-O-β-d-glucoside of trillin, and the 1-O-glycosides of ophiopogonin D (steroidal saponin). In addition, PGase-1 also could hydrolyze the α-d-galactopyranoside, β-d-glucopyranoside, and β-d-galactopyranoside of p-nitrophenyl-glycosides, but the enzyme could not hydrolyze the α-d-mannopyranoside, α-l-arabinopyranoside, α-d-glucopyranoside, β-d-xylopyranoside, and α-l-rhamnopyranoside of p-nitrophenyl-glycosides. These new properties of PGase-1 were significantly different from those of previously described steroidal saponin-glycosidases and the glycosidases currently described in Enzyme Nomenclature by the NC-IUBMB. The gene (termed as pgase-1) encoding PGase-1 was cloned, sequenced, and expressed in Pichia pastoris GS115. The complete nucleotide sequence of pgase-1 consists of 1,725 bp. The recombinant PGase-1 from recombinant P. pastoris GS115 strain also showed the activity hydrolyzing glycosides of steroidal saponins which was similar to that of the wild-type PGase-1 from A. oryzae. The PGase-1 gene is highly similar to Aspergilli α-amylase (EC 3.2.1.1), and PGase-1 should be classified as glycoside hydrolase family 13 by the method of gene sequence-based classification. But the enzyme properties of PGase-1 are different from those of α-amylase in this family.  相似文献   

6.
In continuation of our search for new antimicrobial secondary metabolites from Bacillus cereus associated with rhabditid entomopathogenic nematode, a new microbial diketopiperazine, cyclo(l-Pro-d-Arg), was isolated from the ethyl acetate extract of fermented modified nutrient broth. The chemical structures of the isolated compounds were identified based on their 1D, 2D NMR and high-resolution electrospray ionisation–mass spectroscopy data. Antibacterial activity of the compound was determined by minimum inhibitory concentration and disc diffusion method against medically important bacteria, and the compound was recorded to have significant antibacterial activity against test bacteria. The highest activity was recorded against Klebsiella pneumoniae (1 μg/mL). Cyclo(l-Pro-d-Arg) was recorded to have significant antitumor activity against HeLa cells (IC50 value 50 μg/mL), and this compound was recorded to have no cytotoxicity against normal monkey kidney cells (VERO) up to 100 μg/mL). To the best of our knowledge, this is the first time that cyclo(l-Pro-d-Arg) has been isolated from a microbial natural source.  相似文献   

7.
Cationic antimicrobial peptides (AMPs) have attracted a great interest as novel class of antibiotics that might help in the treatment of infectious diseases caused by pathogenic bacteria. However, some AMPs with high antimicrobial activities are also highly hemolytic and subject to proteolytic degradation from human and bacterial proteases that limit their pharmaceutical uses. In this work a d-diastereomer of Pandinin 2, d-Pin2, was constructed to observe if it maintained antimicrobial activity in the same range as the parental one, but with the purpose of reducing its hemolytic activity to human erythrocytes and improving its ability to resist proteolytic cleavage. Although, the hydrophobic and secondary structure characteristics of l- and d-Pin2 were to some extent similar, an important reduction in d-Pin2 hemolytic activity (30–40 %) was achieved compared to that of l-Pin2 over human erythrocytes. Furthermore, d-Pin2 had an antimicrobial activity with a MIC value of 12.5 μM towards Staphylococcus aureus, Escherichia coli, Streptococcus agalactiae and two strains of Pseudomonas aeruginosa in agar diffusion assays, but it was half less potent than that of l-Pin2. Nevertheless, the antimicrobial activity of d-Pin2 was equally effective as that of l-Pin2 in microdilution assays. Yet, when d- and l-Pin2 were incubated with trypsin, elastase and whole human serum, only d-Pin2 kept its antimicrobial activity towards all bacteria, but in diluted human serum, l- and d-Pin2 maintained similar peptide stability. Finally, when l- and d-Pin2 were incubated with proteases from P. aeruginosa DFU3 culture, a clinical isolated strain, d-Pin2 kept its antibiotic activity while l-Pin2 was not effective.  相似文献   

8.
Twelve different amino acids were each substituted for threonine-654 in a cloned glucansucrase from Leuconostoc mesenteroides NRRL B-1118. Both the native and the cloned enzyme with threonine at position 654 produced a water-insoluble glucan containing approximately 44 mol% 1,3-disubstituted α-d-glucopyranosyl units and 29 mol% 1,6-disubstituted α-d-glucopyranosyl units. Several substitutions yielded an enzyme that produced an increased percentage of 1,3-disubstituted α-d-glucopyranosyl units, with corresponding decreases in 1,6-disubstituted α-d-glucopyranosyl units. Only one substitution, tyrosine, resulted in a significant increase in the percentage of 1,6-disubstituted α-d-glucopyranosyl units, with a concomitant increase in glucan yield. The mutated enzymes that produced the highest levels of 1,3-disubstituted α-d-glucopyranosyl units were also significantly activated by the addition of dextran, but glucan yields were also lower in these mutants.  相似文献   

9.
A gene in Bradyrhizobium japonicum USDA 110, annotated as a ribitol dehydrogenase (RDH), had 87 % sequence identity (97 % positives) to the N-terminal 31 amino acids of an l-glucitol dehydrogenase from Stenotrophomonas maltophilia DSMZ 14322. The 729-bp long RDH gene coded for a protein consisting of 242 amino acids with a molecular mass of 26.1 kDa. The heterologously expressed protein not only exhibited the main enantio selective activity with d-glucitol oxidation to d-fructose but also converted l-glucitol to d-sorbose with enzymatic cofactor regeneration and a yield of 90 %. The temperature stability and the apparent K m value for l-glucitol oxidation let the enzyme appear as a promising subject for further improvement by enzyme evolution. We propose to rename the enzyme from the annotated RDH gene (locus tag bll6662) from B. japonicum USDA as a d-sorbitol dehydrogenase (EC 1.1.1.14).  相似文献   

10.
It has long been believed that amino acids comprising proteins of all living organisms are only of the l-configuration, except for Gly. However, peptidyl d-amino acids were observed in hydrolysates of soluble high molecular weight fractions extracted from cells or tissues of various organisms. This strongly suggests that significant amounts of d-amino acids are naturally present in usual proteins. Thus we analyzed the d-amino acid contents of His-tag-purified β-galactosidase and human urocortin, which were synthesized by Escherichia coli grown in controlled synthetic media. After acidic hydrolysis for various times at 110°C, samples were derivatized with 4-fluoro-7-nitro-2, 1, 3-benzoxadiazole (NBD-F) and separated on a reverse-phase column followed by a chiral column into d- and l-enantiomers. The contents of d-enantiomers of Ala, Leu, Phe, Val, Asp, and Glu were determined by plotting index d/(d + l) against the incubation time for hydrolysis and extrapolating the linear regression line to 0 h to eliminate the effect of racemization of amino acids during the incubation. Significant contents of d-amino acids were reproducibly detected, the d-amino acid profile being specific to an individual protein. This finding indicated the likelihood that d-amino acids are in fact present in the purified proteins. On the other hand, the d-amino acid contents of proteins were hardly influenced by the addition of d- or l-amino acids to the cultivation medium, whereas intracellular free d-amino acids sensitively varied according to the extracellular conditions. The origin of these d-amino acids detected in proteins was discussed.  相似文献   

11.
Four potential dehydrogenases identified through literature and bioinformatic searches were tested for l-arabonate production from l-arabinose in the yeast Saccharomyces cerevisiae. The most efficient enzyme, annotated as a d-galactose 1-dehydrogenase from the pea root nodule bacterium Rhizobium leguminosarum bv. trifolii, was purified from S. cerevisiae as a homodimeric protein and characterised. We named the enzyme as a l-arabinose/d-galactose 1-dehydrogenase (EC 1.1.1.-), Rl AraDH. It belongs to the Gfo/Idh/MocA protein family, prefers NADP+ but uses also NAD+ as a cofactor, and showed highest catalytic efficiency (k cat/K m) towards l-arabinose, d-galactose and d-fucose. Based on nuclear magnetic resonance (NMR) and modelling studies, the enzyme prefers the α-pyranose form of l-arabinose, and the stable oxidation product detected is l-arabino-1,4-lactone which can, however, open slowly at neutral pH to a linear l-arabonate form. The pH optimum for the enzyme was pH 9, but use of a yeast-in-vivo-like buffer at pH 6.8 indicated that good catalytic efficiency could still be expected in vivo. Expression of the Rl AraDH dehydrogenase in S. cerevisiae, together with the galactose permease Gal2 for l-arabinose uptake, resulted in production of 18 g of l-arabonate per litre, at a rate of 248 mg of l-arabonate per litre per hour, with 86 % of the provided l-arabinose converted to l-arabonate. Expression of a lactonase-encoding gene from Caulobacter crescentus was not necessary for l-arabonate production in yeast.  相似文献   

12.
Selected eight yeast strains isomerized-glycero-d-galacto-heptose tod-manno-heptulose. The conversion is 7–10%. Under identical conditions, the reverse isomerization ofd-manno-heptulose tod-glycero-d-galacto-heptose ord-glycero-d-talo-heptose does not take place.  相似文献   

13.
We identified ergothionase, which catalyzes conversion of ergothioneine to thiolurocanic acid and trimethylamine, in a newly isolated ergothioneine-utilizing strain, Burkholderia sp. HME13. The enzyme was purified and its N-terminal amino acid sequence was determined. Based on the amino acid sequence, the gene encoding the enzyme was cloned and expressed in Escherichia coli. The recombinant enzyme was purified to homogeneity and characterized. The enzyme consisted of four identical 55-kDa subunits. The enzyme showed maximum activity at pH 8.0 and 65 °C and was stable between pH 7.0 and pH 10.0 and up to 60 °C. The enzyme acted on ergothioneine (K m: 19 μM, V max: 270 μmol/min/mg), but not d-histidine, l-histidine, d-tyrosine, l-tyrosine, d-phenylalanine, or l-phenylalanine. The enzyme was activated by BaCl2 and strongly inhibited by CuSO4, ZnSO4, and HgCl2. The amino acid sequence of ergothionase showed 23 % similarity to histidine ammonia-lyase (HAL) from Pseudomonas putida and 17 % similarity to phenylalanine ammonia-lyase (PAL) from parsley. However, the tripeptide sequence, Ala-Ser-Gly, which is important for catalysis in both HAL and PAL, was not conserved in ergothionase. The application of ergothionase for the quantification of ergothioneine contained in practical food and blood samples was investigated by performing a recovery test. Satisfactory recovery data (98.7–104 %) were obtained when ergothioneine was added to extract of tamogitake and hemolysis blood.  相似文献   

14.
3-O-β-d-Xylopyranosyl-l-serine (xylosylserine) was synthesized by the following three-step procedure: 1) 2,3,4-tri-O-benzoyl-α-d-xylopyranosyl bromide (benzobromoxylose) was condensed withN-carbobenzoxy-l-serine benzyl ester using the silver triflate-collidine complex as promoter; 2) theN-carbobenzoxy and benzyl ester groups in the resultant glycoside were cleaved by transfer hydrogenation with palladium black as catalyst and ammonium formate as hydrogen donor; and 3) the benzoyl groups were removed with methanolic ammonia. Xylosylserine was obtained in an overall yield of 70%. O-β-d-Galactopyranosyl-(1-4)-O-β-d-xylopyranosyl-(1-3)-l-serine (galactosylxylosylserine) was also synthesized by this methodology and was characterized by 2-dimensional (2D) NMR spectroscopy techniques. The two serine glycosides (xylosylserine and galactosylxylosylserine) were used in detection and partial purification of galactosyltransferase I (UDP-d-galactose:d-xylose galactosyltransferase) from adult rat liver.  相似文献   

15.
A recombinant putative N-acyl-d-glucosamine 2-epimerase from Dictyoglomus turgidum was identified as a cellobiose 2-epimerase by evaluating its substrate specificity. The purified enzyme was a 46?kDa monomer with a specific activity of 16.8?μmol?min?1?mg?1 for cellobiose. The epimerization activity was maximal at pH 7.0 and 70?°C with a half-life of 55?h. The isomerization of the glucose at the reducing end of β-1,4- and α-1,4-linked gluco-oligosaccharides to a fructose moiety by the enzyme took place after the epimerization of the glucose to a mannose moiety. The enzyme converted cellobiose to 12.8?% 4-O-β-d-glucopyranosyl-d-mannose and 54.6?% 4-O-β-d-glucopyranosyl-d-fructose as an equilibrium and converted lactose to 12.8?% epilactose and 54.3?% lactulose.  相似文献   

16.
The demand for d-2-phenylglycine used to synthesize semisynthetic antibiotics and pesticides is increasing. We have isolated a Chryseobacterium sp. that selectively transformed the l-form of racemic d,l-2-phenylglycine to (2S)-2-acetylamide-2-phenylacetic acid with a molar yield of 50 % and an enantiomer excess of >99.5 % under optimal culture conditions, consequently resulting in 99 % pure d-2-phenylglycine remaining in the culture. The enantioselective N-acetylation was catalyzed by an acetyl-CoA-dependent N-acetyltransferase whose synthesis was induced by l-2-phenylglycine. The enzyme differed from previously reported bacterial arylamine N-acetyltransferases in molecular mass and substrate specificity. The relative activity ratio of the enzyme with the substrates l-2-phenylglycine, d-2-phenylglycine, 2-(2-chlorophenyl)glycine, and 5-aminosalicylic acid (a good substrate of arylamine N-acetyltransferase) was 100:0:56.9:5.49, respectively. The biotransformation by the N-acetyltransferase-producing bacterium reported here could constitute a new preparative route for the enzymatic resolution of d,l-2-phenylglycine.  相似文献   

17.
β-d-Xylosidase catalyzes hydrolysis of xylooligosaccharides to d-xylose residues. The enzyme, SXA from Selenomonas ruminantium, is the most active catalyst known for the reaction; however, its activity is inhibited by d-xylose and d-glucose (K i values of ~10?2?M). Higher K i’s could enhance enzyme performance in lignocellulose saccharification processes for bioethanol production. We report here the development of a two-tier high-throughput screen where the 1° screen selects for activity (active/inactive screen) and the 2° screen selects for a higher K i(d-xylose) and its subsequent use in screening ~5,900 members of an SXA enzyme library prepared using error-prone PCR. In one variant, termed SXA-C3, K i(d-xylose) is threefold and K i(d-glucose) is twofold that of wild-type SXA. C3 contains four amino acid mutations, and one of these, W145G, is responsible for most of the lost affinity for the monosaccharides. Experiments that probe the active site with ligands that bind only to subsite ?1 or subsite +1 indicate that the changed affinity stems from changed affinity for d-xylose in subsite +1 and not in subsite ?1 of the two-subsite active site. Trp145 is 6 Å from the active site, and its side chain contacts three active-site residues, two in subsite +1 and one in subsite ?1.  相似文献   

18.
d-Tagatose 3-epimerase family enzymes can efficiently catalyze the epimerization of free keto-sugars, which could be used for d-psicose production from d-fructose. In previous studies, all optimum pH values of these enzymes were found to be alkaline. In this study, a d-psicose 3-epimerase (DPEase) with neutral pH optimum from Clostridium bolteae (ATCC BAA-613) was identified and characterized. The gene encoding the recombinant DPEase was cloned and expressed in Escherichia coli. In order to characterize the catalytic properties, the recombinant DPEase was purified to electrophoretic homogeneity using nickel-affinity chromatography. Ethylenediaminetetraacetic acid was shown to inhibit the enzyme activity completely; therefore, the enzyme was identified as a metalloprotein that exhibited the highest activity in the presence of Co2+. Although the DPEase demonstrated the most activity at a pH ranging from 6.5 to 7.5, it exhibited optimal activity at pH 7.0. The optimal temperature for the recombinant DPEase was 55 °C, and the half-life was 156 min at 55 °C. Using d-psicose as the substrate, the apparent K m, k cat, and catalytic efficiency (k cat/K m) were 27.4 mM, 49 s?1, and 1.78 s?1 mM?1, respectively. Under the optimal conditions, the equilibrium ratio of d-fructose to d-psicose was 69:31. For high production of d-psicose, 216 g/L d-psicose could be produced with 28.8 % turnover yield at pH 6.5 and 55 °C. The recombinant DPEase exhibited weak-acid stability and thermostability and had a high affinity and turnover for the substrate d-fructose, indicating that the enzyme was a potential d-psicose producer for industrial production.  相似文献   

19.
The gene coding for d-psicose 3-epimerase (DPEase) from Clostridium sp. BNL1100 was cloned and expressed in Escherichia coli. The recombinant enzyme was purified by Ni-affinity chromatography. It was a metal-dependent enzyme and required Co2+ as optimum cofactor. It displayed catalytic activity maximally at pH 8.0 and 65 °C (as measured over 5 min). The optimum substrate was d-psicose, and the K m, turnover number (k cat), and catalytic efficiency (k cat/K m) for d-psicose were 227 mM, 32,185 min?1, and 141 min?1 mM?1, respectively. At pH 8.0 and 55 °C, 120 g d-psicose l?1 was produced from 500 g d-fructose l?1 after 5 h.  相似文献   

20.
Ethylene glycol (EG) is an important platform chemical with steadily expanding global demand. Its commercial production is currently limited to fossil resources; no biosynthesis route has been delineated. Herein, a biosynthesis route for EG production from d-xylose is reported. This route consists of four steps: d-xylose?→?d-xylonate?→?2-dehydro-3-deoxy-d-pentonate?→?glycoaldehyde?→?EG. Respective enzymes, d-xylose dehydrogenase, d-xylonate dehydratase, 2-dehydro-3-deoxy-d-pentonate aldolase, and glycoaldehyde reductase, were assembled. The route was implemented in a metabolically engineered Escherichia coli, in which the d-xylose?→?d-xylulose reaction was prevented by disrupting the d-xylose isomerase gene. The most efficient construct produced 11.7 g?L?1 of EG from 40.0 g?L?1 of d-xylose. Glycolate is a carbon-competing by-product during EG production in E. coli; blockage of glycoaldehyde?→?glycolate reaction was also performed by disrupting the gene encoding aldehyde dehydrogenase, but from this approach, EG productivity was not improved but rather led to d-xylonate accumulation. To channel more carbon flux towards EG than the glycolate pathway, further systematic metabolic engineering and fermentation optimization studies are still required to improve EG productivity.  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