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.
A thermostable homodimeric isocitrate dehydrogenase from the hyperthermophilic sulfate-reducing archaeon Archaeoglobus fulgidus was purified and characterized. The mol. mass of the isocitrate dehydrogenase subunit was 42 kDa as determined by SDS-PAGE. Following separation by SDS-PAGE, A. fulgidus isocitrate dehydrogenase could be renatured and detected in situ by activity staining. The enzyme showed dual coenzyme specificity with a high preference for NADP+. Optimal temperature for activity was 90° C or above, and a half-life of 22 min was found for the enzyme when incubated at 90° C in a 50 mM Tricine-KOH buffer (pH 8.0). Based on the N-terminal amino acid sequence, the gene encoding the isocitrate dehydrogenase was cloned. DNA sequencing identified the icd gene as an open reading frame encoding a protein of 412 amino acids with a molecular mass corresponding to that determined for the purified enzyme. The deduced amino acid sequence closely resembled that of the isocitrate dehydrogenase from the archaeon Caldococcus noboribetus (59% identity) and bacterial isocitrate dehydrogenases, with 57% identity with isocitrate dehydrogenase from Escherichia coli. All the amino acid residues directly contacting substrate and coenzyme (except Ile-320) in E. coli isocitrate dehydrogenase are conserved in the enzyme from A. fulgidus. The primary structure of A. fulgidus isocitrate dehydrogenase confirmes the presence of Bacteria-type isocitrate dehydrogenases among Archaea. Multiple alignment of all the available amino acid sequences of di- and multimeric isocitrate dehydrogenases from the three domains of life shows that they can be divided into three distinct phylogenetic groups. Received: 6 February 1997 / Accepted: 12 June 1997  相似文献   

2.
NADP+-specific glutamate dehydrogenase (EC 1.4.1.4) was purified to homogeneity from the extremely thermophilic, strictly anaerobic, sulfate-reducing archaeon Archaeoglobus fulgidus strain 7324. The native enzyme (263 kDa) is composed of subunits of mol. mass 46 kDa, suggesting a hexameric structure. The temperature optimum for enzyme activity was > 95° C. The enzyme was highly thermostable, having a half-life of 140 min at 100° C. Potassium phosphate, KCl, and NaCl enhanced the thermal stability and increased the rate of activity three- to fourfold. The N-terminal 26-amino-acid sequence showed a high degree of similarity to glutamate dehydrogenases from Pyrococcus spp. and Thermococcus spp. Received: 25 March 1997 / Accepted: 11 July 1997  相似文献   

3.
Bacteria that produced NAD+-dependent phenylalanine dehydrogenase (EC 1.4.1.20) were selected among l-methionine utilizers isolated from soil. A bacterial strain showing phenylalanine dehydrogenase activity was chosen and classified in the genus Microbacterium. Phenylalanine dehydrogenase was purified from the crude extract of Microbacterium sp. strain DM 86-1 (TPU 3592) to homogeneity as judged by SDS-polyacrylamide disc gel electrophoresis. The enzyme has an isoelectric point of 5.8 and a relative molecular weight (M r) of approximately 330,000. The enzyme is composed of eight identical subunits with an M r of approximately 41,000. The apparent K m values for l-phenylalanine and NAD+ were calculated to be 0.10 mM and 0.20 mM, respectively. No loss of the enzyme activity was observed upon incubation at 55° C for 10 min. Received: 30 July 1997 / Accepted: 4 November 1997  相似文献   

4.
Methanobacterium thermoautotrophicum (strain Marburg) was found to contain two malate dehydrogenases, which were partially purified and characterized. One was specific for NAD+ and catalyzed the dehydrogenation of malate at approximately one-third of the rate of oxalacetate reduction, and the other could equally well use NAD+ and NADP+ as coenzyme and catalyzed essentially only the reduction of oxalacetate. Via the N-terminal amino acid sequences, the encoding genes were identified in the genome of M. thermoautotrophicum (strain ΔH). Comparison of the deduced amino acid sequences revealed that the two malate dehydrogenases are phylogenetically only distantly related. The NAD+-specific malate dehydrogenase showed high sequence similarity to l-malate dehydrogenase from Methanothermus fervidus, and the NAD(P)+-using malate dehyrogenase showed high sequence similarity to l-lactate dehydrogenase from Thermotoga maritima and l-malate dehydrogenase from Bacillus subtilis. A function of the two malate dehydrogenases in NADPH:NAD+ transhydrogenation is discussed. Received: 29 December 1997 / Accepted: 4 March 1998  相似文献   

5.
A newly isolated gram-negative bacterium, possibly Brevundimonas diminuta, utilised d,l-vanillylmandelate (d,l-VMA) as a sole carbon and energy source. The organism converted d,l-VMA to vanillylglyoxylate using a soluble NAD-dependent dehydrogenase specific for d-VMA and a dye-linked, membrane-associated l-VMA dehydrogenase. Vanillylglyoxylate was further metabolised by decarboxylation, dehydrogenation and demethylation to protocatechuate. A 4,5-dioxygenase cleaved protocatechuate to 2-hydroxy-4-carboxymuconic semialdehyde. Partially purified d-VMA dehydrogenase exhibited optimal activity at 30° C and pH 9.5 and had an apparent K m for d-VMA of 470 μM. Although induced by several substituted mandelates, the enzyme had a narrow substrate specificity range with virtually no activity towards d-mandelate. Such properties render the enzyme of potential use in both diagnostic and biosynthetic applications. Received: 23 January 1996 / Accepted: 9 April 1996  相似文献   

6.
NADP-dependent glutamate dehydrogenase (l-glutamate: NADP oxidoreductase, deaminating, EC 1.4.1.4) from the aerobic hyperthermophilic archaeon Aeropyrum pernix K1 (JCM 9820) was purified to homogeneity for characterization. The enzyme retained its full activity on heating at 95°C for 30 min, and the maximum activity in l-glutamate deamination was obtained around 100°C. The enzyme showed a strict specificity for l-glutamate and NADP on oxidative deamination and for 2-oxoglutarate and NADPH on reductive amination. The K m values for NADP, l-glutamate, NADPH, 2-oxoglutarate, and ammonia were 0.039, 3.3, 0.022, 1.7, and 83 mM, respectively. On the basis of the N-terminal amino acid sequence, the encoding gene was identified in the A. pernix K1 genome, cloned, and expressed in Escherichia coli. Analysis of the nucleotide sequence revealed an open reading frame of 1257 bp starting with a minor TTG codon and encoding a protein of 418 amino acids with a molecular weight of 46 170. Phylogenetic analysis revealed that the glutamate dehydrogenase from A. pernix K1 clustered with those from aerobic Sulfolobus solfataricus, Sulfolobus shibatae, and anaerobic Pyrobaculum islandicum in Crenarchaeota, and it separated from another cluster of the enzyme from Thermococcales in Euryarchaeota. The branching pattern of the enzymes from A. pernix K1, S. solfataricus, S. shibatae, and Pb. islandicum in the phylogenetic tree coincided with that of 16S rDNAs obtained from the same organisms. Received: April 24, 2000 / Accepted: August 10, 2000  相似文献   

7.
Whole-genome sequence analysis of Bacillus halodurans ATCC BAA-125 revealed an isomerase gene (rhaA) encoding an l-rhamnose isomerase (l-RhI). The identified l -RhI gene was cloned from B. halodurans and over-expressed in Escherichia coli. DNA sequence analysis revealed an open reading frame of 1,257 bp capable of encoding a polypeptide of 418 amino acid residues with a molecular mass of 48,178 Da. The molecular mass of the purified enzyme was estimated to be ∼48 kDa by sodium dodecyl sulfate-polyacrylamide gel electrophoresis and 121 kDa by gel filtration chromatography, suggesting that the enzyme is a homodimer. The enzyme had an optimal pH and temperature of 7 and 70°C, respectively, with a k cat of 8,971 min−1 and a k cat/K m of 17 min−1 mM−1 for l-rhamnose. Although l-RhIs have been characterized from several other sources, B. halodurans l-RhI is distinguished from other l-RhIs by its high temperature optimum (70°C) with high thermal stability of showing 100% activity for 10 h at 60°C. The half-life of the enzyme was more than 900 min and ∼25 min at 60°C and 70°C, respectively, making B. halodurans l-RhI a good choice for industrial applications. This work describes one of the most thermostable l-RhI characterized thus far.  相似文献   

8.
α-l-Rhamnosidase was extracted and purified from the cells of Pseudomonas paucimobilis FP2001 with a 19.5% yield. The purified enzyme, which was homogeneous as shown by SDS-PAGE and isoelectric focusing, had a molecular weight of 112,000 and an isoelectric point of 7.1. The enzyme activity was accelerated by Ca2+ and remained stable for several months when stored at –20 °C. The optimum pH was 7.8; the optimum temperature was 45 °C. The K m, V max and k cat for p-nitrophenyl α-l-rhamnopyranoside were 1.18 mM, 92.4 μM · min–1 and 117,000 · min–1, respectively. Examination of the substrate specificity using various synthetic and natural l-rhamnosyl glycosides showed that this enzyme had a relatively broader substrate specificity than those reported so far. Received: 24 May 1999 / Accepted: 7 October 1999  相似文献   

9.
A malate dehydrogenase (MDH) from Streptomyces avermitilis MA-4680 (SaMDH) has been expressed and purified as a fusion protein. The molecular mass of SaMDH is about 35 kDa determined by SDS-PAGE. The recombinant SaMDH has a maximum activity at pH 8.0. The enzyme shows the optimal temperature around 42°C and displays a half-life (t 1/2) of 160 min at 50°C which is more thermostable than reported MDHs from most bacteria and fungi. The k cat value of SaMDH is about 240-fold of that for malate oxidation. In addition, the k cat/K m ratio shows that SaMDH has about 1,246-fold preference for oxaloacetate (OAA) reduction over l-malate oxidation. The recombinant SaMDH may also use NADPH as a cofactor although it is a highly NAD(H)-specific enzyme. There was no activity detected when malate and NADP+ were used as substrates. Substrate inhibition studies show that SaMDH activity is strongly inhibited by excess OAA with NADH, but is not sensitive to excess l-malate. Enzymatic activity is enhanced by the addition of Na+, NH4 +, Ca2+, Cu2+ and Mg2+ and inhibited by addition of Hg2+ and Zn2+. MDH is widely used in coenzyme regeneration, antigen immunoassays and bioreactors. The enzymatic analysis could provide the important basic knowledge for its utilizations.  相似文献   

10.
Haloacid dehalogenases have potential applications in the pharmaceutical and fine chemical industry as well as in the remediation of contaminated land. The l-2-haloacid dehalogenase from the thermophilic archaeon Sulfolobus tokodaii has been cloned and over-expressed in Escherichia coli and successfully purified to homogeneity. Here we report the structure of the recombinant dehalogenase solved by molecular replacement in two different crystal forms. The enzyme is a homodimer with each monomer being composed of a core-domain of a β-sheet bundle surrounded by α-helices and an α-helical sub-domain. This fold is similar to previously solved mesophilic l-haloacid dehalogenase structures. The monoclinic crystal form contains a putative inhibitor l-lactate in the active site. The enzyme displays haloacid dehalogenase activity towards carboxylic acids with the halide attached at the C2 position with the highest activity towards chloropropionic acid. The enzyme is thermostable with maximum activity at 60°C and a half-life of over 1 h at 70°C. The enzyme is relatively stable to solvents with 25% activity lost when incubated for 1 h in 20% v/v DMSO.  相似文献   

11.
Industrially, the use of high temperatures (40–60°C) in the l-malate production process could result in rapid inactivation of the mesophilic fumarases, warranting constant replenishment of the biocatalyst. Thus, a thermostable fumarase C that is active and stable at high temperatures would be ideal. Biochemical studies using recombinant fumarase C from thermophilic Streptomyces thermovulgaris (stFUMC) indicated that it was optimally active at 50°C and highly stable even after 24 h of incubation at 40°C. The same gene from mesophilic Streptomyces coelicolor (scfumC) was also cloned and expressed as soluble proteins for comparison in thermal properties of both enzymes. In contrast to stFUMC, scFUMC exhibited a lower temperature optima of 30°C and was rapidly denatured at 50°C. The specific activity of stFUMC was also higher than that of scFUMC by 20-fold. After primary sequence comparison, three hydrophilic amino acid residues, R163, E170 and S347, were forged into the thermolabile scFUMC either singly or in combination for the investigation of their contributions in the thermal properties of the mutant enzymes. Of the mutants studied, the A347S scFUMC mutant resulted in the highest increase in optimum temperature of 10°C and a fourfold enhancement in specific activity. G163R/G170E and G163R/G170E/A347S scFUMC mutants are more thermostable than wild-type scFUMC. These findings support stFUMC as a highly efficient, thermostable fumarase C with industrial potential and suggest that R163, E170 and S347 are involved in the enhancement of thermal properties in fumarase C.  相似文献   

12.
Thermotoga hypogea is an extremely thermophilic anaerobic bacterium capable of growing at 90°C. It uses carbohydrates and peptides as carbon and energy sources to produce acetate, CO2, H2, l-alanine and ethanol as end products. Alcohol dehydrogenase activity was found to be present in the soluble fraction of T. hypogea. The alcohol dehydrogenase was purified to homogeneity, which appeared to be a homodimer with a subunit molecular mass of 40 ± 1 kDa revealed by SDS-PAGE analyses. A fully active enzyme contained iron of 1.02 ± 0.06 g-atoms/subunit. It was oxygen sensitive; however, loss of enzyme activity by exposure to oxygen could be recovered by incubation with dithiothreitol and Fe2+. The enzyme was thermostable with a half-life of about 10 h at 70°C, and its catalytic activity increased along with the rise of temperature up to 95°C. Optimal pH values for production and oxidation of alcohol were 8.0 and 11.0, respectively. The enzyme had a broad specificity to use primary alcohols and aldehydes as substrates. Apparent K m values for ethanol and 1-butanol were much higher than that of acetaldehyde and butyraldehyde. It was concluded that the physiological role of this enzyme is likely to catalyze the reduction of aldehydes to alcohols.  相似文献   

13.
Galdieria sulphuraria (Galdieri) Merola can grow heterotrophically on at least ten different polyols. We investigated their metabolic path to glycolysis/gluconeogenesis and identified two NAD-dependent polyol dehydrogenases. Activity of other enzymes metabolizing mannitol or sorbitol could not be detected. The two dehydrogenases had a broad substrate specificity and were termed xylitol dehydrogenase (EC 1.1.1.14; substrate specificity: xylitol > d-sorbitol > d-mannitol > l-arabitol) and d-arabitol dehydrogenase (EC 1.1.1.11; substrate specificity: d-arabitol > l-fucitol > d-mannitol > d-threitol) according to the substrate with the lowest K m value. The xylitol dehydrogenase was stable during purification. In contrast, the d-arabitol dehydrogenase was thermolabile and depended on divalent ions for stability and activity, preferentially Mn2+ and Ni2+. The molecular mass of the xylitol dehydrogenase was estimated to be 295 kDa by size-exclusion chromatography and 220 kDa by rate-sedimentation centrifugation. The d-arabitol dehydrogenase had a molecular mass of 105 kDa as determined by rate-sedimentation centrifugation. The specific activity of both enzymes increased about fourfold when cells were transferred from autotrophic to heterotrophic conditions regardless of whether sugars or polyols were supplied as substrates. The significance of polyol metabolism in Galdieria sulphuraria with regard to the natural habitat of the alga is discussed. Received: 15 January 1997 / Accepted: 12 February 1997  相似文献   

14.
β-Poly(l-malate) (PMLA) production in Physarum polycephalum has been followed by using d-[1-13C]glucose and Ca13CO3. Nuclear magnetic resonance studies of PMLA showed that the 13C label from [1-13C]glucose was incorporated in the presence of CaCO3 into positions C-3 (-CH2-) and C-4 (-CO-) of the l-malate repeating unit of PMLA. The 13C label from Ca13CO3 was incorporated into position C-4 and indicated that not only the endogenous CO2 but also the exogenous CO2 from CaCO3 served significantly as a carbon source for PMLA production. In the absence of CaCO3, the 13C labeling pattern of PMLA from d-[1-13C]glucose was almost indistinguishable from that for the natural abundance 13C-NMR spectrum of the polymer. These results indicated that l-malate used for PMLA production is synthesized either via carboxylation of pyruvate and reduction of oxaloacetate in the presence of CaCO3 or via the oxidative tricarboxylic acid (TCA) cycle in the absence of CaCO3. Avidin strongly inhibited the formation of l-malate via carboxylation; the 13C labeling pattern of PMLA in the presence of CaCO3 was almost identical with that for the natural abundance spectrum when avidin was added, indicating that l-malate utilized for PMLA production was supplied under this condition by the oxidative TCA cycle. Received: 16 March 1999 / Received revision: 5 May 1999 / Accepted: 7 May 1999  相似文献   

15.
An α-l-rhamnosidase was purified by fractionating a culture filtrate of Aspergillus kawachii grown on l-rhamnose as the sole carbon source. The α-l-rhamnosidase had a molecular mass of 90 kDa and a high degree of N-glycosylation of approximately 22%. The enzyme exhibited optimal activity at pH 4.0 and temperature of 50 °C. Further, it was observed to be thermostable, and it retained more than 80% of its original activity following incubation at 60 °C for 1 h. Its T 50 value was determined to be 72 °C. The enzyme was able to hydrolyze α-1,2- and α-1,6-glycosidic bonds. The specific activity of the enzyme was higher toward naringin than toward hesperidin. The A. kawachii α-l-rhamnosidase-encoding gene (Ak-rhaA) codes for a 655-amino-acid protein. Based on the amino acid sequence deduced from the cDNA, the protein possessed 13 potential N-glycosylation recognition sites and exhibited a high degree of sequence identity (up to 75%) with the α-l-rhamnosidases belonging to the glycoside hydrolase family 78 from Aspergillus aculeatus and with hypothetical Aspergillus oryzae and Aspergillus fumigatus proteins. Electronic supplementary material  The online version of this article (doi:) contains supplementary material, which is available to authorized users.  相似文献   

16.
We studied the effects of high temperatures and elevated hydrostatic pressures on the physiological behavior and viability of the extremely thermophilic deep-sea archaeon Thermococcus peptonophilus. Maximal growth rates were observed at 30 and 45 MPa although no significant increases in cell yields were detected. Growth at 60 MPa was slower. The optimal growth temperature shifted from 85° C at 30 MPa to 90–95° C at 45 MPa. Cell viability during the stationary phase was also enhanced under high pressure. A trend towards barophily at pressures greater than those encountered in situ at the sea floor was demonstrated at increasing growth temperatures. The viability of cells during starvation, at high temperature (90, 95° C), and at low temperature (10° C) was enhanced at 30 and 45 MPa as compared to atmospheric pressure. These results show that the extremely thermophilic archaeon T. peptonophilus is a barophile. Received: 21 October 1996 / Accepted: 5 February 1997  相似文献   

17.
The gene encoding an α-l-arabinofuranosidase from Geobacillus caldoxylolyticus TK4, AbfATK4, was isolated, cloned, and sequenced. The deduced protein had a molecular mass of about 58 kDa, and analysis of its amino acid sequence revealed significant homology and conservation of different catalytic residues with α-l-arabinofuranosidases belonging to family 51 of the glycoside hydrolases. A histidine tag was introduced at the N-terminal end of AbfATK4, and the recombinant protein was expressed in Escherichia coli BL21, under control of isopropyl-β-D-thiogalactopyranoside-inducible T7 promoter. The enzyme was purified by nickel affinity chromatography. The molecular mass of the native protein, as determined by gel filtration, was about 236 kDa, suggesting a homotetrameric structure. AbfATK4 was active at a broad pH range (pH 5.0–10.0) and at a broad temperature range (40–85°C), and it had an optimum pH of 6.0 and an optimum temperature of 75–80°C. The enzyme was more thermostable than previously described arabinofuranosidases and did not lose any activity after 48 h incubation at 70°C. The protein exhibited a high level of activity with p-nitrophenyl-α-l-arabinofuranoside, with apparent K m and V max values of 0.17 mM and 588.2 U/mg, respectively. AbfATK4 also exhibited a low level of activity with p-nitrophenyl-β-d-xylopyranoside, with apparent K m and V max values of 1.57 mM and 151.5 U/mg, respectively. AbfATK4 released l-arabinose only from arabinan and arabinooligosaccharides. No endoarabinanase activity was detected. These findings suggest that AbfATK4 is an exo-acting enzyme.  相似文献   

18.
A laccase from Pycnoporus sanguineus was purified by two steps using phenyl-Sepharose columm. A typical procedure provided 54.1-fold purification, with a yield of 8.37%, using syringaldazine as substrate. The molecular weight of the purified laccase was 69 and 68 kDa as estimated by 12% (w/v) SDS-PAGE gel and by gel filtration, respectively. The K m values for the substrates ABTS, syringaldazine, and guaiacol were 58, 8.3, and 370 μM, respectively. The enzyme’s pH optimum for syringaldazine was 4.2 and optimal activity was 50°C. The enzyme showed to be thermostable because when kept at 50°C for 24 and 48 h it retained 93 and 76% activity. This laccase was inhibited by l-cysteine, β-mercaptoethanol, NaN3, NaF, and HgCl2.  相似文献   

19.
Fumarase from the syntrophic propionate-oxidizing bacterium strain MPOB was purified 130-fold under anoxic conditions. The native enzyme had an apparent molecular mass of 114 kDa and was composed of two subunits of 60 kDa. The enzyme exhibited maximum activity at pH 8.5 and approximately 54° C. The K m values for fumarate and l-malate were 0.25 mM and 2.38 mM, respectively. Fumarase was inactivated by oxygen, but the activity could be restored by addition of Fe2+ and β-mercaptoethanol under anoxic conditions. EPR spectroscopy of the purified enzyme revealed the presence of a [3Fe-4S] cluster. Under reducing conditions, only a trace amount of a [4Fe-4S] cluster was detected. Addition of fumarate resulted in a significant increase of this [4Fe-4S] signal. The N-terminal amino acid sequence showed similarity to the sequences of fumarase A and B of Escherichia coli (56%) and fumarase A of Salmonella typhimurium (63%). Received: 15 September 1995 / Accepted: 13 November 1995  相似文献   

20.
The aldehyde dehydrogenase activity of the sulfate-reducing bacterium Desulfovibrio simplex strain DSM 4141 was characterized in cell-free extracts. Oxygen-sensitive, constitutive aldehyde dehydrogenase activity was found in cells grown on l(+)-lactate, hydrogen, or vanillin with sulfate as the electron acceptor. A 1.83- to 2.6-fold higher specific activity was obtained in cells grown in media supplemented with 1 μM WO4 2–. The aldehyde dehydrogenase in cell-free extracts catalyzed the oxidation of aliphatic (K m < 20 μM) and aromatic aldehydes (K m < 0.32 mM) using methyl viologen as the electron acceptor. Flavins (FMN and FAD) were also active and are proposed to be the natural cofactors, while no activity was obtained with NAD+ or NADP+. 185WO4 2– was incorporated in vivo into D. simplex; it was found exclusively in the soluble fraction (≥ 98%). Anionic-exchange chromatography demonstrated coelution of 185W with two distinct peaks, the first one containing hydrogenase and formate dehydrogenase activities, and the second one aldehyde dehydrogenase activity. Received: 7 February 1997 / Accepted: 6 June 1997  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