共查询到20条相似文献,搜索用时 15 毫秒
1.
Rik I. L. Eggen Ans C. M. Geerling Wilfried G. B. Voorhorst Remco Kort Willem M. de Vos 《Biocatalysis and Biotransformation》1994,11(2):131-141
The gene for glutamate dehydrogenase (GDH) from Pyrococcus furiosus has been cloned, sequenced and expressed in Escherichia coli. Significant GDH activity could be detected in this host, allowing the further structure-function analysis of this hyperthermostable hexameric enzyme. The deduced primary sequence of the P. furiosus GDH was homologous to various bacterial, archaeal and eukaryal GDHs. Detailed comparative analysis of the primary sequences of these GDHs suggest that a decrease in Gly residues can be a general stabilizing feature of proteins functional under extreme conditions such as high temperatures or high salt concentrations whereas an increase in He residues and a decrease in Cys residues is typical for hyperthermostable enzymes. 相似文献
2.
Hiroshi Nishise Susumu Maehashi Hideaki Yamada Yoshiki Tani 《Bioscience, biotechnology, and biochemistry》2013,77(12):3347-3353
The addition of an organic solvent to the reaction mixture led to a change in the activity of glycerol dehydrogenase. The activity on glycerol decreased to lower than that on 1,2-propanediol. This was due to the apparent increase in the Km value for each substrate. The system was applied to determination of 1,2-propanediol. The standard curve for 1,2-propanediol with a rate assay method was not affected by a 10-fold amount of glycerol in the presence of «-butanol. Chemical modification, except for succinylation of the NH2-residue, did not cause any change in substrate specificity. The participation of a His-residue in the active site was suggested by the results of chemical modification. 相似文献
3.
Ohnishi M Saito M Wakabayashi S Ishizuka M Nishimura K Nagata Y Kasai S 《Journal of bacteriology》2008,190(4):1359-1365
Pyrobaculum islandicum is an anaerobic hyperthermophilic archaeon that is most active at 100 degrees C. A pyridoxal 5'-phosphate-dependent serine racemase called Srr was purified from the organism. The corresponding srr gene was cloned, and recombinant Srr was purified from Escherichia coli. It showed the highest racemase activity toward L-serine, followed by L-threonine, D-serine, and D-threonine. Like rodent and plant serine racemases, Srr is bifunctional, showing high L-serine/L-threonine dehydratase activity. The sequence of Srr is 87% similar to that of Pyrobaculum aerophilum IlvA (a putative threonine dehydratase) but less than 32% similar to any other serine racemases and threonine dehydratases. Sodium dodecyl sulfate-polyacrylamide gel electrophoresis and gel filtration analyses revealed that Srr is a homotrimer of a 44,000-molecular-weight subunit. Both racemase and dehydratase activities were highest at 95 degrees C, while racemization and dehydration were maximum at pH 8.2 and 7.8, respectively. Unlike other, related Ilv enzymes, Srr showed no allosteric properties: neither of these enzymatic activities was affected by either L-amino acids (isoleucine and valine) or most of the metal ions. Only Fe2+ and Cu2+ caused 20 to 30% inhibition and 30 to 40% stimulation of both enzyme activities, respectively. ATP inhibited racemase activity by 10 to 20%. The Km and Vmax values of the racemase activity of Srr for L-serine were 185 mM and 20.1 micromol/min/mg, respectively, while the corresponding values of the dehydratase activity of L-serine were 2.2 mM and 80.4 micromol/min/mg, respectively. 相似文献
4.
Satomura T Kusumi K Ohshima T Sakuraba H 《Bioscience, biotechnology, and biochemistry》2011,75(10):2049-2051
A gene encoding a UDP-glucose dehydrogenase homologue was identified in the hyperthermophilic archaeon, Pyrobaculum islandicum. This gene was expressed in Escherichia coli, and the product was purified and characterized. The expressed enzyme is the most thermostable UDP-glucose dehydrogenase so far described, with a half-life of 10 min at 90 °C. The enzyme retained its full activity after incubating in a pH range of 5.0-10.0 for 10 min at 80 °C. The temperature dependence of the kinetic parameters for this enzyme was examined at 37-70 °C. A decrease in K(m)s for UDP-glucose and NAD was observed with decreasing temperature. This resulted in the enzyme still retaining high catalytic efficiency (V(max)/K(m)) for the substrate and cofactor, even at 37 °C. These characteristics make the enzyme potentially useful for its application at a much lower temperature such as 37 °C than the optimum growth temperature of 100 °C for P. islandicum. 相似文献
5.
Purification and Properties of NAD-Dependent Sorbitol Dehydrogenase from Apple Fruit 总被引:7,自引:0,他引:7
This is the first report of the purification of NAD-dependentsorbitol dehydrogenase (NAD-SDH) from a plant source. The enzymewas extracted from apple (Malus domestica cv. Ourin) fruit andpurified until it appeared as a single polypeptide chain ona gel after SDS-PAGE. From the apparent molecular mass of 62kDa obtained by SDS-PAGE and that of 120 kDa by gel filtration,the enzyme appeared to be a homodimer. Maximum rates of oxidationof sorbitol and reduction of fructose were observed at pH 9.6and pH 6.0, respectively. The Km for oxidation of sorbitol was40.3 mM and that for reduction of fructose was 215 mM. The maximumrate of oxidation of sorbitol was about 10 times higher thanthat of the reduction of fructose. The results of the kineticanalysis strongly suggest that in vivo the enzyme would favorthe conversion of sorbitol to fructose over the reverse reaction.None of the divalent cations tested had any effect on the oxidationof sorbitol by NAD-SDH. The reaction catalyzed by NAD-SDH wasnot specific to sorbitol and other substrates could also beoxidized. Among the tested substrates, ethyl alcohol had a particularlyhigh affinity for the enzyme. (Received February 2, 1994; Accepted May 31, 1994) 相似文献
6.
7.
Cloning and characterization of a family B DNA polymerase from the hyperthermophilic crenarchaeon Pyrobaculum islandicum
下载免费PDF全文

In order to extend the limited knowledge about crenarchaeal DNA polymerases, we cloned a gene encoding a family B DNA polymerase from the hyperthermophilic crenarchaeon Pyrobaculum islandicum. The enzyme shared highest sequence identities with a group of phylogenetically related DNA polymerases, designated B3 DNA polymerases, from members of the kingdom Crenarchaeota, Pyrodictium occultum and Aeropyrum pernix, and several members of the kingdom Euryarchaeota. Six highly conserved regions as well as a DNA-binding motif, indicative of family B DNA polymerases, were identified within the sequence. Furthermore, three highly conserved 3'-5' exonuclease motifs were also found. The gene was expressed in Escherichia coli, and the DNA polymerase was purified to homogeneity by heat treatment and affinity chromatography. Activity staining after sodium dodecyl sulfate-polyacrylamide gel electrophoresis revealed an active polypeptide of approximately 90 kDa. For the recombinant DNA polymerase from P. islandicum, activated calf thymus DNA was used as a substrate rather than primed single-stranded DNA. The enzyme was strongly inhibited by monovalent cations and N-ethylmaleimide; it is moderately sensitive to aphidicolin and dideoxyribonucleoside triphosphates. The half-life of the enzyme at 100 and 90 degrees C was 35 min and >5 h, respectively. Interestingly, the pH of the assay buffer had a significant influence on the 3'-5' exonuclease activity of the recombinant enzyme. Under suitable assay conditions for PCR, the enzyme was able to amplify lambda DNA fragments of up to 1,500 bp. 相似文献
8.
Agathe Garnier Ali Berredjem Bernard Botton 《Fungal genetics and biology : FG & B》1997,22(3):168-176
The NAD-dependent glutamate dehydrogenase (GDH) (EC 1.4.1.2) fromLaccaria bicolorwas purified 410-fold to apparent electrophoretic homogeneity with a 40% recovery through a three-step procedure involving ammonium sulfate precipitation, anion-exchange chromatography on DEAE–Trisacryl, and gel filtration. The molecular weight of the native enzyme determined by gel filtration was 470 kDa, whereas sodium dodecyl sulfate–polyacrylamide gel electrophoresis gave rise to a single band of 116 kDa, suggesting that the enzyme is composed of four identical subunits. The enzyme was specific for NAD(H). The pH optima were 7.4 and 8.8 for the amination and deamination reactions, respectively. The enzyme was found to be highly unstable, with virtually no activity after 20 days at −75°C, 4 days at 4°C, and 1 h at 50°C. The addition of ammonium sulfate improved greatly the stability of the enzyme and full activity was still observed after several months at −75°C. NAD-GDH activity was stimulated by Ca2+and Mg2+but strongly inhibited by Cu2+and slightly by the nucleotides AMP, ADP, and ATP. The Michaelis constants for NAD, NADH, 2-oxoglutarate, and ammonium were 282 μM, 89 μM, 1.35 mM, and 37 mM, respectively. The enzyme had a negative cooperativity for glutamate (Hill number of 0.3), and itsKmvalue increased from 0.24 to 3.6 mM when the glutamate concentration exceeded 1 mM. These affinity constants of the substrates, compared with those of the NADP-GDH of the fungus, suggest that the NAD-GDH is mainly involved in the catabolism of glutamate, while the NADP-GDH is involved in the catalysis of this amino acid. 相似文献
9.
10.
Goda S Kojima M Nishikawa Y Kujo C Kawakami R Kuramitsu S Sakuraba H Hiragi Y Ohshima T 《Biochemistry》2005,44(46):15304-15313
The specific activity of recombinant Pyrobaculum islandicum glutamate dehydrogenase (pis-GDH) expressed in Escherichia coli is much lower than that of the native enzyme. However, when the recombinant enzyme is heated at 90 degrees C or exposed to 5 M urea, the activity increases to a level comparable to that of the native enzyme. Small-angle X-ray scattering measurements revealed that the radius of gyration (R(g,z)) of the hexameric recombinant enzyme was reduced to 47 A from 55 A by either heat or urea, and that the final structure of the active enzyme is the same irrespective of the mechanism of activation. Activation was accompanied by a shift in the peaks of the Kratky plot, though the molecular mass of the enzyme was unchanged. The activation-induced decline in R(g,z) followed first-order kinetics, indicating that activation of the enzyme involved a transition between two states, which was confirmed by singular-value decomposition analysis. When the low-resolution structure of the recombinant enzyme was restored using ab initio modeling, we found it to possess no point symmetry, whereas the heat-activated enzyme possessed 32-point symmetry. In addition, a marked increase in the fluorescence emission was observed with addition of ANS to the inactive recombinant enzyme but not the active forms, indicating that upon activation hydrophobic residues on the surface of the recombinant protein moved to the interior. Taken together, these data strongly suggest that subunit rearrangement, i.e., a change in the quaternary structure of the hexameric recombinant pis-GDH, is essential for activation of the enzyme. 相似文献
11.
Tatiana N. Stekhanova Andrey V. Mardanov Ekaterina Y. Bezsudnova Vadim M. Gumerov Nikolai V. Ravin Konstantin G. Skryabin Vladimir O. Popov 《Applied and environmental microbiology》2010,76(12):4096-4098
Short-chain alcohol dehydrogenase, encoded by the gene Tsib_0319 from the hyperthermophilic archaeon Thermococcus sibiricus, was expressed in Escherichia coli, purified and characterized as an NADPH-dependent enantioselective oxidoreductase with broad substrate specificity. The enzyme exhibits extremely high thermophilicity, thermostability, and tolerance to organic solvents and salts.Alcohol dehydrogenases (ADHs; EC 1.1.1.1.) catalyze the interconversion of alcohols to their corresponding aldehydes or ketones by using different redox-mediating cofactors. NAD(P)-dependent ADHs, due to their broad substrate specificity and enantioselectivity, have attracted particular attention as catalysts in industrial processes (5). However, mesophilic ADHs are unstable at high temperatures, sensitive to organic solvents, and often lose activity during immobilization. In this relation, there is a considerable interest in ADHs from extremophilic microorganisms; among them, Archaea are of great interest. The representatives of all groups of NAD(P)-dependent ADHs have been detected in genomes of Archaea (11, 12); however, only a few enzymes have been characterized, and the great majority of them belong to medium-chain (3, 4, 14, 16, 19) or long-chain iron-activated ADHs (1, 8, 9). Up to now, a single short-chain archaeal ADH from Pyrococcus furiosus (10, 18) and only one archaeal aldo-keto reductase also from P. furiosus (11) have been characterized.Thermococcus sibiricus is a hyperthermophilic anaerobic archaeon isolated from a high-temperature oil reservoir capable of growth on complex organic substrates (15). The complete genome sequence of T. sibiricus has been recently determined and annotated (13). Several ADHs are encoded by the T. sibiricus genome, including three short-chain ADHs (Tsib_0319, Tsib_0703, and Tsib_1998) (13). In this report, we describe the cloning and expression of the Tsib_0319 gene from T. sibiricus and the purification and the biochemical characterization of its product, the thermostable short-chain ADH (TsAdh319).The Tsib_0319 gene encodes a protein with a size of 234 amino acids and the calculated molecular mass of 26.2 kDa. TsAdh319 has an 85% degree of sequence identity with short-chain ADH from P. furiosus (AdhA; PF_0074) (18). Besides AdhA, close homologs of TsAdh319 were found among different bacterial ADHs, but not archaeal ADHs. The gene flanked by the XhoI and BamHI sites was PCR amplified using two primers (sense primer, 5′-GTTCTCGAGATGAAGGTTGCTGTGATAACAGGG-3′, and antisense primer, 5′-GCTGGATCCTCAGTATTCTGGTCTCTGGTAGACGG-3′) and cloned into the pET-15b vector. TsAdh319 was overexpressed, with an N-terminal His6 tag in Escherichia coli Rosetta-gami (DE3) and purified to homogeneity by metallochelating chromatography (Hi-Trap chelating HP column; GE Healthcare) followed by gel filtration on Superdex 200 10/300 GL column (GE Healthcare) equilibrated in 50 mM Tris-HCl (pH 7.5) with 200 mM NaCl. The homogeneity and the correspondence to the calculated molecular mass of 28.7 kDa were verified by SDS-PAGE (7). The molecular mass of native TsAdh319 was 56 to 60 kDa, which confirmed the dimeric structure in solution.The standard ADH activity measurement was made spectrophotometrically at the optimal pH by following either the reduction of NADP (in 50 mM Gly-NaOH buffer; pH 10.5) or the oxidation of NADPH (in 0.1 M sodium phosphate buffer; pH 7.5) at 340 nm at 60°C. The enzyme exhibited a strong preference for NADP(H) and broad substrate specificity (Table (Table1).1). The highest oxidation rates were found with pentoses d-arabinose (2.0 U mg−1) and d-xylose (2.46 U mg−1), and the highest reduction rates were found with dimethylglyoxal (5.9 U mg−1) and pyruvaldehyde (2.2 U mg−1). The enzyme did not reduce sugars which were good substrates for the oxidation reaction. The kinetic parameters of TsAdh319 determined for the preferred substrates are shown in Table Table2.2. The enantioselectivity of the enzyme was estimated by measuring the conversion rates of 2-butanol enantiomers. TsAdh319 showed an evident preference, >2-fold, for (S)-2-butanol over (RS)-2-butanol. The enzyme stereoselectivity is confirmed by the preferred oxidation of d-arabinose over l-arabinose (Table (Table1).1). The fact that TsAdh319 is metal independent was supported by the absence of a significant effect of TsAdh319 preincubation with 10 mM Me2+ for 30 min before measuring the activity in the presence of 1 mM Me2+ or EDTA (Table (Table3).3). TsAdh319 also exhibited a halophilic property, so the enzyme activity increased in the presence of NaCl and KCl and the activation was maintained even at concentration of 4 M and 3 M, respectively (Table (Table33).
Open in a separate windowaSubstrates were present in 250 mM or 50 mM (*) concentrations.bRelative rates, measured under standard conditions, were calculated by defining the activity for 2-propanol as 100%, which corresponds to 1.0 U mg−1. Data are averages from triplicate experiments.cRelative rates, measured under standard conditions, were calculated by defining the activity for pyruvaldehyde as 100%, which corresponds to 2.2 U mg−1. Data are averages from triplicate experiments.
Open in a separate windowaActivity was measured under standard conditions with 2-propanol. Data are averages from triplicate experiments.bActivity was measured under standard conditions with pyruvaldehyde. Data are averages from triplicate experiments.
Open in a separate windowaThe activity was measured under standard conditions with 2-propanol; relative rates were calculated by defining the activity without salts as 100%, which corresponds to 0.9 U mg−1. Data are averages from duplicate experiments.The most essential distinctions of TsAdh319 are the thermophilicity and high thermostability of the enzyme. The optimum temperature for the 2-propanol oxidation catalyzed by TsAdh319 was not achieved. The initial reaction rate of oxidation increased up to 100°C (Fig. (Fig.1).1). The Arrhenius plot is a straight line, typical of a single rate-limited thermally activated process, but there is no obvious transition point due to the temperature-dependent conformational changes of the protein molecule. The activation energy for the oxidation of 2-propanol was estimated at 84.0 ± 5.8 kJ·mol−1. The thermostability of TsAdh319 was calculated from residual TsAdh319 activity after preincubation of 0.4 mg/ml enzyme solution in 50 mM Tris-HCl buffer (pH 7.5) containing 200 mM NaCl at 70, 80, 90, or 100°C. The preincubation at 70°C or 80°C for 1.5 h did not cause a decrease in the TsAdh319 activity, but provoked slight activation. The residual TsAdh319 activities began to decrease after 2 h of preincubation at 70°C or 80°C and were 10% and 15% down from the control, respectively. The determined half-life values of TsAdh319 were 2 h at 90°C and 1 h at 100°C.Open in a separate windowFIG. 1.Temperature dependence of the initial rate of the 2-propanol reduction by TsAdh319. The reaction was initiated by enzyme addition to a prewarmed 2-propanol-NADP mixture. The inset shows the Arrhenius plot of the same data.Protein thermostability often correlates with such important biotechnological properties as increased solvent tolerance (2). We tested the influence of organic solvents at a high concentration (50% [vol/vol]) on TsAdh319 by using either preincubation of the enzyme at a concentration of 0.2 mg/ml with solvents for 4 h at 55°C or solvent addition into the reaction mixture to distinguish the effect of solvent on the protein stability and on the enzyme activity. TsAdh319 showed significant solvent tolerance in both cases (Table (Table4),4), and the effects of solvents could be modulated by salts, acting apparently as molecular lyoprotectants (17). Furthermore, TsAdh319 maintained 57% of its activity in 25% (vol/vol) 2-propanol, which could be used as the cosubstrate in cofactor regeneration (6).
Open in a separate windowaThe activity measured at the standard condition with 2-propanol as a substrate. Data are averages from triplicate experiments.bPreincubation for 4 h at 55°C in the presence of 50% (vol/vol) of solvent prior the activity assay.cWithout preincubation, solvent addition to the reaction mixture up to 50% (vol/vol) or using the buffer saturated by a solvent (*).dDMSO, dimethyl sulfoxide.eDMFA, dimethylformamide.From all the aforesaid we may suppose TsAdh319 or its improved variant to be interesting both for the investigation of structural features of protein tolerance and for biotechnological applications. 相似文献
TABLE 1.
Substrate specificity of TsAdh319Substratea | Relative activity (%) |
---|---|
Oxidation reactionb | |
Methanol | 0 |
2-Methoxyethanol | 0 |
Ethanol | 36 |
1-Butanol | 80 |
2-Propanol | 100 |
(RS)-(±)-2-Butanol | 86 |
(S)-(+)-2-Butanol | 196 |
2-Pentanol | 67 |
1-Phenylmethanol | 180 |
1.3-Butanediol | 91 |
Ethyleneglycol | 0 |
Glycerol | 16 |
d-Arabinose* | 200 |
l-Arabinose* | 17 |
d-Xylose* | 246 |
d-Ribose* | 35 |
d-Glucose* | 146 |
d-Mannose* | 48 |
d-Galactose* | 0 |
Cellobiose* | 71 |
Reduction reactionc | |
Pyruvaldehyde | 100 |
Dimethylglyoxal | 270 |
Glyoxylic acid | 36 |
Acetone | 0 |
Cyclopentanone | 0 |
Cyclohexanone | 4 |
3-Methyl-2-pentanone* | 13 |
d-Arabinose* | 0 |
d-Xylose* | 0 |
d-Glucose* | 0 |
Cellobiose* | 0 |
TABLE 2.
Apparent Km and Vmax values for TsAdh319Coenzyme or substrate | Apparent Km (mM) | Vmax (U mg−1) | kcat (s−1) |
---|---|---|---|
NADPa | 0.022 ± 0.002 | 0.94 ± 0.02 | 0.45 ± 0.01 |
NADPHb | 0.020 ± 0.003 | 3.16 ± 0.11 | 1.51 ± 0.05 |
2-Propanol | 168 ± 29 | 1.10 ± 0.09 | 0.53 ± 0.04 |
d-Xylose | 54.4 ± 7.4 | 1.47 ± 0.09 | 0.70 ± 0.04 |
Pyruvaldehyde | 17.75 ± 3.38 | 4.26 ± 0.40 | 2.04 ± 0.19 |
TABLE 3.
Effect of various ions and EDTA on TsAdh319aCompound | Concn (mM) | Relative activity (%) |
---|---|---|
None | 0 | 100 |
NaCl | 400 | 206 |
600 | 227 | |
4,000 | 230 | |
KCl | 600 | 147 |
2,000 | 200 | |
3,000 | 194 | |
MgCl2 | 10 | 78 |
CoCl2 | 10 | 105 |
NiSO4 | 10 | 100 |
ZnSO4 | 10 | 79 |
FeSO4 | 10 | 74 |
EDTA | 1 | 100 |
5 | 80 |
TABLE 4.
Influence of various solvents on TsAdh319 activityaSolvent | Relative activity (%)b | Relative activity (%)c | |
---|---|---|---|
Buffer without NaCl | Buffer with 600 mM NaCl | ||
None | 100 | 100 | 100 |
DMSOd | 98 | 0 | 40 |
DMFAe | 101 | 13 | 41 |
Methanol | 98 | 25 | 9 |
Acetonitrile | 95 | 0 | 0 |
Ethyl acetate | 47 | 0* | 33* |
Chloroform | 105 | 79* | 81* |
n-Hexane | 105 | 60* | 118* |
n-Decane | 36 | 91* | 107* |
12.
Carmen Pire Frutos C. Marhuenda-egea Julia Esclapez Luis Alcaraz Juan Ferrer Maria Jos Bonete 《Biocatalysis and Biotransformation》2004,22(1):17-23
Reverse micelles were used as a cytoplasmic model to study the kinetics of an extreme halophilic enzyme such as the recombinant glucose dehydrogenase from the Archaeon Haloferax mediterranei. This enzyme was solubilized in reverse micelles of hexadecyltrimethylammoniumbromide in cyclohexane, with 1-butanol as co-surfactant. Glucose dehydrogenase retained its catalytic properties in this organic medium, showing good stability at low water content, even at low salt concentration (125 mM NaCl). The dependence of the enzymatic activity on the molar water surfactant ratio (w0=[H2O]/[surfactant]) increased with rising water content. Surprisingly, the activity of this extreme halophilic enzyme did not depend on the salt concentration in reverse micelles. The kinetic of the enzymatic oxidation of β-D-glucose to D-glucono-1,5-lactone using NADP+ as coenzyme for the glucose dehydrogenase from Haloferax mediterranei was also studied in the reverse micellar system. 相似文献
13.
An Unusual Oxygen-Sensitive, Iron- and Zinc-Containing Alcohol Dehydrogenase from the Hyperthermophilic Archaeon Pyrococcus furiosus
下载免费PDF全文

Pyrococcus furiosus is a hyperthermophilic archaeon that grows optimally at 100°C by the fermentation of peptides and carbohydrates to produce acetate, CO2, and H2, together with minor amounts of ethanol. The organism also generates H2S in the presence of elemental sulfur (S0). Cell extracts contained NADP-dependent alcohol dehydrogenase activity (0.2 to 0.5 U/mg) with ethanol as the substrate, the specific activity of which was comparable in cells grown with and without S0. The enzyme was purified by multistep column chromatography. It has a subunit molecular weight of 48,000 ± 1,000, appears to be a homohexamer, and contains iron (~1.0 g-atom/subunit) and zinc (~1.0 g-atom/subunit) as determined by chemical analysis and plasma emission spectroscopy. Neither other metals nor acid-labile sulfur was detected. Analysis using electron paramagnetic resonance spectroscopy indicated that the iron was present as low-spin Fe(II). The enzyme is oxygen sensitive and has a half-life in air of about 1 h at 23°C. It is stable under anaerobic conditions even at high temperature, with half-lives at 85 and 95°C of 160 and 7 h, respectively. The optimum pH for ethanol oxidation was between 9.4 and 10.2 (at 80°C), and the apparent Kms (at 80°C) for ethanol, acetaldehyde, NADP, and NAD were 29.4, 0.17, 0.071, and 20 mM, respectively. P. furiosus alcohol dehydrogenase utilizes a range of alcohols and aldehydes, including ethanol, 2-phenylethanol, tryptophol, 1,3-propanediol, acetaldehyde, phenylacetaldehyde, and methyl glyoxal. Kinetic analyses indicated a marked preference for catalyzing aldehyde reduction with NADPH as the electron donor. Accordingly, the proposed physiological role of this unusual alcohol dehydrogenase is in the production of alcohols. This reaction simultaneously disposes of excess reducing equivalents and removes toxic aldehydes, both of which are products of fermentation. 相似文献
14.
谷氨酸脱氢酶 (GDH)是谷氨酸生物合成的关键酶 ,谷氨酸棒杆菌S91 1 4是目前我国味精工业应用最广泛的生产菌种 ,其谷氨酸脱氢酶的研究尚未见报道。分离纯化该菌中的谷氨酸脱氢酶 ,研究其辅酶组成 ,对揭示谷氨酸脱氢酶的分子结构和性质 ,提高谷氨酸产率很有必要。将培养至对数期中期的细胞离心收集并用含适量DTT、ED TA的Tris_HCl缓冲液 (pH 7 5 )洗涤 ,用Frenchpressurecellpress破碎 ,离心去除菌体碎片得无细胞抽提液。然后使用 KTA_10 0快速纯化系统经DEAE_纤维素柱、疏水柱 (HIC)、G_2 0 0凝胶过滤柱层析得到纯化大约 70倍的以NAD PH为辅酶的GDH和部分纯化的以NADH辅酶的GDH。这两个酶分别对NADPH、NADH高度专一 ,不能相互代替。经HPLC和SDS_PAGE测得前一种酶的分子量和亚基分子量分别为 188kD和 32kD ,表明该酶为具有相同亚基的六聚体。酶活性测定使用HITACHIU_30 0 0分光光度计利用NAD(P)H在 340nm氧化的初速度进行。蛋白质含量测定利用Bradford方法进行 ,并以牛血清白蛋白为标准蛋白。纯化结果表明S91 1 4中确实存在两种GDH ,其中以NADH为辅酶的GDH尚未见报道。和某些具有两种GDH的微生物一样 ,S91 1 4可能也是以NADPH为辅酶的GDH参与谷氨酸的合成代谢 ,以NADH为辅酶的GDH参与谷氨酸的分解代谢。 相似文献
15.
Tsutomu Yamaguchi 《Bioscience, biotechnology, and biochemistry》2013,77(12):3363-3368
Most of the organic solvents tested were found to stimulate the mutagenicities of some mutagenic compounds. These stimulated compounds were restricted to the sugar-degradation products among various mutagenic compounds. From reduction of Nitro Blue tetrazolium (NBT), it can be said that all of the sugar-degradation compounds tested formed oxygen radicals in alkali conditions by autoxidation, and it was confirmed that the rates of the reducion of NBT by these compounds were greatly stimulated by various organic solvents. Further, the depolymerization of DNA by sugar-degradation compounds was found to stimulated by organic solvents. These results demonstrate that organic solvents enhance the oxygen radical formation of these sugar-degradation compounds, which leads to the stimulation of their mutagenicity. 相似文献
16.
Effects of Polar Organic Solvents on the Biocatalysis of Chlorophyllase in a Biphasic Organic System
《Bioscience, biotechnology, and biochemistry》2013,77(11):1947-1952
The effects of polarity of various organic solvents, including acetone, ethanol, and propanol, used in a biphasic organic system, on the hydrolytic activity of a partially purified chlorophyllase from Phaeodactylum tricornutum were investigated. The different concentrations of each polar organic solvent, from 0 to 40%, were added to a mixture (45:55, v/v) of hexane and a buffer solution of Tris–HCl (20 mm, pH 7.5). The most appropriate concentrations of acetone, ethanol, and propanol for the hydrolytic activity of chlorophyllase were 12.5, 5.0, and 2.5%, respectively. The results indicated that the optimum reaction time for the chlorophyllase activity in the biphasic system decreased from 7.0 h to 3.0, 5.0, and 5.0 h, respectively, upon the addition of an appropriate amount of acetone, ethanol, or propanol. The Vmax and Km as well as the inhibitory effect of phytol on the chlorophyllase activity in the biphasic organic system containing a polar organic solvent were also investigated. 相似文献
17.
18.
利用简并引物和RT-PCR方法从金针菇(Flammulina velutipes)幼嫩子实体中克隆获得FvGDH全长cDNA序列.构建入门载体pGWC-FvGDH,利用Gateway克隆技术的LR反应构建原核重组表达载体pDESTl7-FvGDH,转化大肠杆菌BL21(DE3).通过IPTG法诱导表达融合蛋白并进行表达条件优化.SDS-PAGE蛋白电泳分析表明,融合蛋白相对分子质量约为53 kD,与预测的一致.最佳表达条件为温度30℃、IPTG浓度0.4mmol/L、诱导4 h.融合蛋白表达量较高,实现了FvGDH的高效表达,并利用Western blotting对其特异性进行鉴定.FvGDH基因高效原核表达体系的成功建立,为进一步研究FvGDH的酶促动力学奠定了基础. 相似文献
19.
The kinetic parameters, kcat and KM, for the hydrolysis of N-α-tosyl-L-arginine methyl ester (1, TAME) by the wild-type subtilisins Carlsberg and BPN′ as well as the BPN′ mutants Glyl66Ser, GLyl66Asn, and Met222Phe, were determined in the presence of 5 and 15% (v/v) of a selection of water-soluble organic solvents. The goals were to compare and evaluate the solvent effects with a view to expanding their use in organic synthetic applications of the WT and mutant subtilisins. The results showed that subtilisin BPN′ and its mutants were much less affected by organic solvents than subtilisin Carlsberg. The BPN′ mutant Met222Phe demonstrated the greatest resistance to cosolvent inactivation, making it a particularly attractive mutant for peptide synthesis. Dimethyl sulfoxide, acetone, and branched alcohols were found to be the most benign solvents, whereas dioxane, THF, and N-methyl-2-pyrrolidinone seriously reduced catalytic activities, even at low concentrations. The results parallel the solvent-effect data available for other proteinases, including α-chymotrypsin. 相似文献
20.
Purification and Characterization of a Soluble and a Particulate Glutamate Dehydrogenase from Rat Brain 总被引:3,自引:4,他引:3
Arlene D. Colon reas Plaitakis Antonis Perakis Soll Berl Donald D. Clarke 《Journal of neurochemistry》1986,46(6):1811-1819
Glutamate dehydrogenase (GDH) activity was determined in high-speed fractions (100,000 g for 60 min) obtained from whole rat brain homogenates after removal of a low-speed pellet (480 g for 10 min). Approximately 60% of the high-speed GDH activity was particulate (associated with membrane) and the remaining was soluble (probably of mitochondrial matrix origin). Most of the particulate GDH activity resisted extraction by several commonly used detergents, high concentration of salt, and sonication; however, it was largely extractable with the cationic detergent cetyltrimethylammonium bromide (CTAB) in hypotonic buffer solution. The two GDH activities were purified using a combination of hydrophobic interaction, ion exchange, and hydroxyapatite chromatography. Throughout these purification steps the two activities showed similar behavior. Kinetic studies indicated similar Km values for the two GDH fractions for the substrates alpha-ketoglutarate, ammonia, and glutamate; however, there were small but significant differences in Km values for NADH and NADPH. Although the allosteric stimulation by ADP and L-leucine and inhibition by diethylstilbestrol was comparable, the two GDH components differed significantly in their susceptibility to GTP inhibition in the presence of 1 mM ADP, with apparent Ki values of 18.5 and 9.0 microM GTP for the soluble and particulate fractions, respectively. The Hill plot coefficient, binding constant, and cooperativity index for the GTP inhibition were also significantly different, indicating that the two GDH activities differ in their allosteric sites. In addition, enzyme activities of the two purified proteins exhibited a significant difference in thermal stability when inactivated at 45 degrees C and pH 7.4 in 50 mM phosphate buffer. 相似文献