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 281 毫秒
1.
Wang X  Fang B  Luo J  Li W  Zhang L 《Biotechnology letters》2007,29(9):1409-1412
The xylose reductase (XR) gene (xyl1) from Candida shehatae was cloned and expressed in Escherichia coli, and purified as a His6-tagged fusion protein. The recombinant XR had Km values for NADH than NADPH of 150 μM and 20 μM, respectively. The optimal reaction was at pH 6.5 and 35°C. The enzyme was specific for d-xylese.  相似文献   

2.
A thermo stable xylanase was purified and characterized from the cladodes of Cereus pterogonus plant species. The enzyme was purified to homogeneity by ammonium sulfate (80%) fractionation, ion exchange and size exclusion chromatography. The enzyme showed a final specific activity of 216.2 U/mg and the molecular mass of the protein was 80 KDa. The optimum pH and temperature for xylanase activity were 5.0 and 80 °C, respectively,. With oat spelt xylan as a substrate the enzyme yielded a Km value of 2.24 mg/mL and a Vmax of 5.8 μmol min−1 mg−1. In the presence of metal ions (1 mM) such as Co2+,Mn2+, Ni2+, Ca2+ and Fe3+ the activity of the enzyme increased, where as strong inhibition of the enzyme activity was observed with the use of Hg2+, Cd2+, Cu2+, while partial inhibition was noted with Zn2+ and Mg2+. The substrate specificity of the xylanase yielded maximum activity with oat spelt xylan.  相似文献   

3.
A psychrophilic glutathione reductase from Antarctic ice microalgae Chlamydomonas sp. Strain ICE-L was purified by ammonium sulfate fractionation and three steps of chromatography. The yield was up to 25.1% of total glutathione reductase in the crude enzyme extract. The glutathione reductase activity was characterized by the spectrophotometric method under different conditions. Purified glutathione reductase was separated by SDS-PAGE, which furnished a homogeneous band. The native molecular mass of the enzyme was 115 kDa. Apparent Km values for NADPH and NADH (both at 0.5 mmol L−1 oxidized glutathione) were 22.3 and 83.8 μmol L−1, respectively. It was optimally active at pH 7.5, and it was stable from pH 5 to 9. Its optimum temperature was 25°C, with activity at 0°C 23.5% of the maximum. Its optimum ion strength and optimum Mg2+ were 50–90 and 7.5 mmol L−1, respectively. Ca2+, Mg2+, and cysteine substantially increased the activity of the enzyme but chelating agents, heavy metals (Cd2+, Pb2+, Cu2+, Zn2+, etc.), NADPH, and ADP had significant inhibitory effects. This glutathione reductase can be used to study the adaptation and mechanism of catalysis of psychrophilic enzymes, and it has a high potential as an environmental biochemical indicator under extreme conditions.  相似文献   

4.
Glutathione reductase [GR, E.C.1.8.1.7] catalyses NADPH dependent reduction of glutathione disulfide (GSSG) to reduced glutathione (GSH). Thus, it is the crucial enzyme to maintain high [GSH]/[GSSG] ratio and physiological redox status in cells. Kidney and liver tissues were considered as a rich source of GR. In this study, rat kidney GR was purified and some of its properties were investigated. The enzyme was purified 2,356 fold with a yield of 16% by using heat-denaturation and Sephadex G25 gel filtration, 2′,5′-ADP Agarose 4B, PBE94 column chromatographies. The purified enzyme had a specific activity (Vm) of 250 U/mg protein and the ratio of absorbances at wavelengths of A 273/A 463, A 280/A 460, A 365/A 460, and A 379/A 463, were 7.1, 6.8, 1.2 and 1.0, respectively. Each mol of GR subunit bound 0.97 mol of FAD. NADH was used as a coenzyme by rat kidney GR but with a lower efficiency (32.7%) than NADPH. Its subunit molecular weight was estimated as 53 kDa. An optimum pH of 6.5 and optimum temperature of 65 °C were found for rat kidney GR. Its activation energy (Ea) and temperature coefficient (Q10) were calculated as 7.02 kcal/mol and 1.42, respectively. The Km(NADPH) and kcat/Km (NADPH) values were found to be 15.3 ± 1.4 μM and 1.68 × 107 M−1 s−1 for the concentration range of 10-200 μM NADPH and when GSSG is the variable substrate, the Km(GSSG) and the kcat/Km(GSSG) values of 53.1 ± 3.4 μM and 4.85 × 106 M−1 s−1 were calculated for the concentration range of 20–1,200 μM GSSG.  相似文献   

5.
Sah S  Phale PS 《Biodegradation》2011,22(3):517-526
1-Naphthol 2-hydroxylase (1-NH) which catalyzes the conversion of 1-naphthol to 1,2-dihydroxynaphthalene was purified to homogeneity from carbaryl-degrading Pseudomonas sp. strain C6. The enzyme was found to be a homodimer with subunit molecular weight of 66 kDa. UV, visible and fluorescence spectral properties, identification of flavin moiety by HPLC as FAD, and reconstitution of apoenzyme by FAD suggest that enzyme is FAD-dependent. 1-NH accepts electron from NADH as well as NADPH. Besides 1-naphthol (K m, 9.1 μM), the enzyme also accepts 5-amino 1-naphthol (K m, 6.4 μM) and 4-chloro 1-naphthol (K m, 2.3 μM) as substrates. Enzyme showed substrate inhibition phenomenon at high concentration of 1-naphthol (K i, 283 μM). Stoichiometric consumption of oxygen and NADH, and biochemical properties suggest that 1-NH belongs to FAD containing external flavomonooxygenase group of oxido-reductase class of enzymes. Based on biochemical and kinetic properties, 1-NH from Pseudomonas sp. strain C6 appears to be different than that reported earlier from Pseudomonas sp. strain C4. Chemical modification and protection by 1-naphthol and NADH suggest that His, Arg, Cys, Tyr and Trp are at or near the active site of 1-NH.  相似文献   

6.
A protocol was developed for rapid and efficient production of transgenic celery plants via somatic embryo regeneration from Agrobacterium tumefaciens- inoculated leaf sections, cotyledons and hypocotyls. These explants were excised from in vitro seedlings of the cvs. XP166 and XP85 and inoculated with A. tumefaciens strain EHA105 containing the binary vector pBISN1. PBISN1 has the neomycin phosphotransferase gene (nptII) and an intron interrupted β-glucuronidase (GUS) reporter gene (gusA). Co-cultivation was carried out for 4 d in the dark on callus induction medium (CIM): Gamborg B5 + 2.79 μM kinetin + 2.26 μM 2,4-dichlorophenoxyacetic acid (2,4-D) supplemented with 100 μM acetosyringone. Embryogenic calluses resistant to kanamycin (Km) were then recovered on CIM + 25 mg l−1 Km + 250 mg l−1 timentin after 12 weeks. Subsequently, a large number of Km-resistant and GUS-positive transformants, tens to hundreds per explant were regenerated via somatic embryogenesis on Gamborg B5 + 4.92 μM 6 (γ,γ-dimethylallylamino)-purine (2iP) + 1.93 μM α-naphthaleneacetic acid (NAA) + 25 mg l−1 Km + 250 mg l−1 timentin after 8 weeks. Using this protocol, the transformation frequency was 5.0% and 5.0% for leaf sections, 17.8% and 18.3% for cotyledons, and 15.9% and 16.7% for hypocotyl explants of cvs. XP85 and XP166, respectively. Stable integration of the model transgenes with 1–3 copy numbers was confirmed in all ten randomly selected transgenic events by Southern blot analysis of gusA. Progeny analysis by histochemical GUS assay showed stable Mendelian inheritance of the transgenes. Thus, A. tumefaciens-mediated transformation of cotyledons or hypocotyls provides an effective and reproducible protocol for large-scale production of transgenic celery plants.  相似文献   

7.
Redox interconversion of glutathione reductase was studiedin situ withS. cerevisiae. The enzyme was more sensitive to redox inactivation in 24 hour-starved cells than in freshly-grown ones. While 5 μM NADPH or 100 μM NADH caused 50% inactivation in normal cells in 30 min, 0.75 μM NADPH or 50 μM NADH promoted a similar effect in starved cells. GSSG reactivated the enzyme previously inactivated by NADPH, ascertaining that the enzyme was subjected to redox interconversion. Low EDTA concentrations fully protected the enzyme from NADPH inactivation, thus confirming the participation of metals in such a process. Extensive inactivation was obtained in permeabilized cells incubated with glucose-6-phosphate or 6-phosphogluconate, in agreement with the very high specific activities of the corresponding dehydrogenases. Some inactivation was also observed with malate, L-lactate, gluconate or isocitrate in the presence of low NADP+ concentrations. The inactivation of yeast glutathione reductase has also been studiedin vivo. The activity decreased to 75% after 2 hours of growth with glucono-δ-lactone as carbon source, while NADPH rose to 144% and NADP+ fell to 86% of their initial values. Greater changes were observed in the presence of 1.5 μM rotenone: enzymatic activity descended to 23% of the control value, while the NADH/NAD+ and NADPH/NADP+ ratios rose to 171% and 262% of their initial values, respectively. Such results indicate that the lowered redox potential of the pyridine nucleotide pool existing when glucono-δ-lactone is oxidized promotesin vivo inactivation of glutathione reductase.  相似文献   

8.
Xylose reductase (XR) is a key enzyme in xylose metabolism because it catalyzes the reduction of xylose to xylitol. In order to study the characteristics of XR from Candida tropicalis SCTCC 300249, its XR gene (xyll) was cloned and expressed in Escherichia coli BL21 (DE3). The fusion protein was purified effectively by Ni2+-chelating chromatography, and the kinetics of the recombinant XR was investigated. The Km values of the C. tropicalis XR for NADPH and NADH were 45.5 μM and 161.9 μM, respectively, which demonstrated that this XR had dual coenzyme specificity. Moreover, this XR showed the highest catalytic efficiency (kcat=1.44×l04 min−1) for xylose among the characterized aldose reductases. Batch fermentation was performed with Saccharomyces serivisiae W303-lA:pYES2XR, and resulted in 7.63 g/L cell mass, 93.67 g/L xylitol, and 2.34 g/L · h xylitol productivity. This XR coupled with its dual coenzyme specificity, high activity, and catalytic efficiency proved its utility in in vitro xylitol production.  相似文献   

9.
1,4-Benzoquinone reductase was purified to electrophoretic homogeneity from bovine liver, and the purified enzyme found to have a molecular mass of 29 kDa, as determined by sodium dodecyl sulfate-polyacrylamide gel electrophoresis. The enzyme exhibited pH optimum between 8.0 and 8.5. The apparentK m for 1,4-benzoquinone was 1.643 mM, and the apparent Km for NADH was 1.837 mM. Various divalent cations, such as Hg2+, Cu2+, and Zn2+, exhibited strong inhibitory effects. The enzyme activity was also strongly inhibited by quercetin, dicumarol, and benzoic acid. Incubation of the enzyme withN-bromosuccinimide and pyridoxal 5′-phosphate led to inhibitions of 100% and 99%, respectively. Accordingly, these results suggest that tryptophan and lysine residues are involved at or near the active sites of the enzyme.  相似文献   

10.
Liu J  Zhang X  Zhou S  Tao P  Liu J 《Current microbiology》2007,54(2):102-107
Chlamydophila pneumoniae AR39 is an obligate intracellular pathogen that causes human acute and chronic respiratory tract diseases. One protein from C. pneumoniae AR39 was assigned as 4-hydroxybenzoate decarboxylase (HBDC). Assays done with the purified oxygen-sensitive protein showed that the optimum pH and temperature were 7.5 and 30°C, respectively. The Km and Vmax obtained for 4-hydroxybenzoate were approximately 0.21 mM and 11.9 nM min−1 mg−1, respectively. During the period of 4-hydroxybenzoate decarboxylation, overall activity of the thermal-sensitive protein was 5.06 nM min−1 mg−1 protein. The 4-hydroxybenzoate decarboxylation was promoted by Mg2+, Fe2+, Mn2+, and Ca2+ but not by Cu2+ or Zn2+. The enzyme also slowly catalyzed the reverse reaction, which was phenol carboxylation.  相似文献   

11.
Bacteroids formed by Mesorhizobium ciceri CC 1192 in symbiosis with chickpea plants (Cicer arietinum L.) contained a single form of citrate synthase [citrate oxaloacetate-lyase (CoA-acetylating) enzyme; EC 4.1.3.7], which had the same electrophoretic mobility as the enzyme from the free-living cells. The citrate synthase from CC 1192 bacteroids had a native molecular mass of 228 ± 32 kDa and was activated by KCl, which also enhanced stability. Double reciprocal plots of initial velocity against acetyl-CoA concentration were linear, whereas the corresponding plots with oxaloacetate were nonlinear. The K m value for acetyl-CoA was 174 μM in the absence of added KCl, and 88 μM when the concentration of KCl in reaction mixtures was 100 mM. The concentrations of oxaloacetate for 50% of maximal activity were 27 μM without added KCl and 14 μM in the presence of 100 mM KCl. Activity of citrate synthase was inhibited 50% by 80 μM NADH and more than 90% by 200 μM NADH. Inhibition by NADH was linear competitive with respect to acetyl-CoA (K is = 23.1 ± 3 μM) and linear noncompetitive with respect to oxaloacetate (K is = 56 ± 3.8 μM and K ii = 115 ± 15.4 μM). NADH inhibition was relieved by NAD+ and by micromolar concentrations of 5′-AMP. In the presence of 50 or 100 mM KCl, inhibition by NADH was apparent only when the proportion of NADH in the nicotinamide adenine dinucleotide pool was greater than 0.6. In the microaerobic environment of bacteroids, NADH may be at concentrations that are inhibitory for citrate synthase. However, this inhibition is likely to be relieved by NAD+ and 5′-AMP, allowing carbon to enter the tricarboxylic acid cycle. Received: 14 July 1999 / Accepted: 20 September 1999  相似文献   

12.
Drew DP  Lunde C  Lahnstein J  Fincher GB 《Planta》2007,225(4):945-954
Monodehydroascorbate reductase (MDHAR; EC 1.6.5.4) catalyses the reduction of the monodehydroascorbate (MDHA) radical to ascorbate, using NADH or NADPH as an electron donor, and is believed to be involved in maintaining the reactive oxygen scavenging capability of plant cells. This key enzyme in the ascorbate-glutathione cycle has been studied here in the moss Physcomitrella patens, which is tolerant to a range of abiotic stresses and is increasingly used as a model plant. In the present study, three cDNAs encoding different MDHAR isoforms of 47 kDa were identified in P. patens, and found to exhibit enzymic characteristics similar to MDHARs in vascular plants despite low-sequence identity and a distant evolutionary relationship between the species. The three cDNAs for the P. patens MDHAR enzymes were expressed in Escherichia coli and the active enzymes were purified and characterized. Each recombinant protein displayed an absorbance spectrum typical of flavoenzymes and contained a single non-covalently bound FAD coenzyme molecule. The K m and k cat values for the heterologously expressed PpMDHAR enzymes ranged from 8 to 18 μM and 120–130 s−1, respectively, using NADH as the electron donor. The K m values were at least an order of magnitude higher for NADPH. The K m values for the MDHA radical were ∼0.5–1.0 μM for each of the purified enzymes, and further kinetic analyses indicated that PpMDHARs follow a ‘ping–pong’ kinetic mechanism. In contrast to previously published data, site-directed mutagenesis indicated that the conserved cysteine residue is not directly involved in the reduction of MDHA.  相似文献   

13.
A ferric reductase was purified into an electrophoretically homologous state from Magnetospirillum gryphiswaldense MSR-1 strain. The enzyme was found within the cytoplasm and associated with the cytoplasmic membrane. The molecular weight of the purified enzyme was calculated as 16.1 kDa using sodium dodecyl sulfate–polyacrylamide gel electrophoresis, and was almost identical to that calibrated using agarose gel filtration chromatography. It was NADH-dependent and required flavin mononucleotide as a cofactor. The optimal reaction temperature and pH values were 30°C and 6.5, respectively. The K m and Vmax values for ferric citrate were 45.1 μM and 1.216 μM min−1, respectively. Though ferric reductase activity could be inhibited by Co2+, Cu2+, Mn2+, and Zn2+, even high concentrations of Mg2+ ions have failed to accomplish such enzyme inhibition. Furthermore, the molecular weight, the N-terminal sequence, and the activity of ferric reductase from MSR-1 are not matching with the enzyme preparation obtained from an analogous strain M. magnetotacticum (MS-1). Therefore, it is concluded that the ferric reductase of M. grysphiwaldense and M. magnetotacticum strains are two different enzymes.  相似文献   

14.
Lu Y  Yu Y  Zhou R  Sun W  Dai C  Wan P  Zhang L  Hao D  Ren H 《Biotechnology letters》2011,33(6):1159-1167
A novel 2,4-dichlorophenol hydroxylase (TfdB, EC 1.14.13.20) gene, designated as tfdB-JLU, was identified from a metagenome constructed from polychlorinated biphenyl-contaminated soil by functional screening and heterologously expressed in Escherichia coli. The deduced amino acid sequence of tfdB-JLU exhibited less than 48% homology with other known TfdBs. The enzyme exhibited a wider substrate spectrum than the previously reported TfdBs and higher relative activity towards ortho-substituted dichlorophenols, 2-chlorophenol, and 3-chlorophenol than towards 2,4-dichlorophenol, the preferred substrate of other known TfdBs. The enzyme had a K m of 5 μM for 2,4-dichlorophenol and 6 μM for NADPH. The optimal temperature and pH of the enzyme were 25°C and 7.5, respectively. Activity of the purified TfdB-JLU was slightly enhanced by Ca2+, Mn2+, Co2+, and Fe2+, and completely inhibited by Cu2+, Hg2+, and Zn2+. This study is the first report to identify a novel TfdB from a metagenome.  相似文献   

15.
The gene encoding carboxylesterase from the hyperthermophilic bacterium Thermotoga maritima (tm0053) was cloned. The recombinant protein (EST53) was overexpressed in Escherichia coli without its NH2-terminal hydrophobic region, and with a C-terminal hexahistidine sequence. The enzyme was purified to homogeneity by heat treatment, followed by Ni2+ affinity chromatography, and then characterized. Among the p-nitrophenyl esters tested, the best substrate was p-nitrophenyl decanoate with K m and k cat values of 3.1 μM and 10.8 s−1, respectively, at 60°C and pH 7.5. The addition of O,O′-bis(2-aminoethyl)ethyleneglycol-N,N,N′,N′-tetraacetic acid decreased the esterase activity, indicating that EST53 is dependent on the presence of Ca2+ ion. In addition, its activity was inhibited by the addition of phenylmethylsulfonyl fluoride and diethyl pyrocarbonate, indicating that it contains serine and histidine residues, which play key roles in the catalytic mechanism. EST53 shows a relatively high degree of similarity to Burkholderia lipases that belong to family I.2 of the lipolytic enzymes, whereas the local sequence around the pentapeptide of EST53 is most similar to those of Bacillus lipases belonging to family I.4.  相似文献   

16.
We have partially purified S-adenosylmethionine decarboxylase (EC 4.1.1.50, SAMDC) from carnation (Dianthus caryophyllus L.) petals and generated polyclonal antibodies against CSDC 16 protein (Leeet al., 1996) overexpressed inE. coli. The protein has been purified approximately 126.8 fold through the steps involving ammonium sulfate fractionation, DEAE-Sepharose column chromatography and Sephacryl S-300 gel filtration. Its molecular mass was 42 kDa in native form and we could also detect a band of 32 kDa molecular mass on SDS-PAGE in western blot analysis using the polyclonal antibodies. The Km value of this enzyme forS-adenosylmethionine was 26.3 μM. The optimum temperature and pH forS-adenosylmethionine decarboxylase activity were 35°C and pH 8.0, respectively. Putrescine and Mg2+ had no effects on the activation of the enzyme activity. Mg2+ did not have any significant effects on the enzyme activity. SAMDC activity was inhibited by putrescine, spermidine and spermine. Methylglyoxal bis-(guanylhydrazone) (MGBG), carbonyl reagents such as hydroxylamine and phenylhydrazine, and sulfhydryl reagent such as 5,5′dithio-bis (2-nitrobenzoic acid) (DTNB) were effective inhibitors of the enzyme. However, isonicotinic acid hydrazide known as an inhibitor of 5′-pyridoxal phosphate (PLP) dependent enzyme activity had no significant effect on the enzyme activity. These results and our previously reported results (Leeet al., 1997b) suggest thatS-adenosylmethionine decarboxylase is a heterodimer, αβ, and some carbonyl group and sulfhydryl group are involved in the catalytic activity.  相似文献   

17.
Igamberdiev AU  Bykova NV  Hill RD 《Planta》2006,223(5):1033-1040
NADH-dependent NO scavenging in barley extracts is linked to hemoglobin (Hb) expression and is inhibited by SH-reagents. Barley Hb has a single cysteine residue. To determine whether this cysteine was critical for NO scavenging, barley Hb and a mutated version, in which the single Cys79 was replaced by Ser, were over-expressed in Escherichia coli and purified to near homogeneity. The purified proteins exhibited very low NO-scavenging activity (12–14 nmol min−1 mg−1 protein) in the presence of NADH or NADPH. This activity was insensitive to SH-reagents. Addition of an extract from barley roots to either of the purified proteins resulted in high NADH-dependent NO turnover in a reaction that was sensitive to SH-reagents. A protein was purified from barley roots and identified by mass-spectrometry analysis as a cytosolic monodehydroascorbate reductase. It efficiently supported NADH-dependent NO scavenging in the presence of either native or mutated barley Hb. Ascorbate strongly facilitated the rate of metHb reduction. The K m for Hb was 0.3 μM, for ascorbate 0.6 mM and for NADH 4 μM. The reaction in the presence of monodehydroascorbate reductase was sensitive to SH-reagents with either form of the Hb. We conclude that metHb reduction and NO turnover do not involve direct participation of the Cys79 residue of barley Hb. NO scavenging is facilitated by monodehydroascorbate reductase mediating a coupled reaction involving ferric Hb reduction in the presence of ascorbate and NADH.  相似文献   

18.
Summary Redox inactivation of glutathione reductase involves metal cations, since chelators protected against NADPH-inactivation, 3 µM EDTA or 10 µM DETAPAC yielding full protection. Ag+, Zn2+ and Cd2+ potentiated the redox inactivation promoted by NADPH alone, while Cr3+, Fe2+, Fe3+, Cu+, and Cu2+ protected the enzyme. The Zn2+ and Cd2+ effect was time-dependent, unlike conventional inhibition. Glutathione reductase interconversion did not require dioxygen, excluding participation of active oxygen species produced by NADPH and metal cations. One Zn2+ ion was required per enzyme subunit to yield full NADPH-inactivation, the enzyme being reactivated by EDTA. Redox inactivation of glutathione reductase could arise from the blocking of the dithiol formed at the active site of the reduced enzyme by metal cations, like Zn2+ or Cd2+.The glutathione reductase activity of yeast cell-free extracts was rapidly inactivated by low NADPH or moderate NADH concentrations; NADP+ also promoted rapid inactivation in fresh extracts, probably after reduction to NADPH. Full inactivation was obtained in cell-free extracts incubated with glucose-6-phosphate or 6-phosphogluconate; the inactivating efficiency of several oxidizable substrates was directly proportional to the specific activities of the corresponding dehydrogenases, confirming that redox inactivation derives from NADPH formed in vitro.Abbreviations DETAPAC diethylenetriaminepentaacetic acid - 2,5-ADP-Sepharose-N6-(6-aminohexyl) adenosine 2,5-bisphosphateSepharose  相似文献   

19.
The G2ALT gene was cloned and sequenced from the thermophilic bacterium Anoxybacillus gonensis G2. The gene is 666 bp long and encodes a protein 221 amino acids in length. The gene was overexpressed in E. coli and purified to homogeneity and biochemically characterized. The enzyme has a molecular mass of 24.5 kDa and it could be classified as a member of the family of bacterial aluminium resistance proteins based on homology searches. When this fragment was expressed in E. coli, it endowed E. coli with Al tolerance to 500 μM. The purified G2ALT protein is active at a broad pH range (pH 4.0–10.0) and temperature range (25°C–80°C) with optima of 6.0 and the apparent optimal temperature of 73°C respectively. Under optimal conditions, G2ALT exhibited a low ATPase activity with K m and V max values of 10±0.55 μM and 26.81±0.13 mg Pi released/min/mg enzyme, respectively. The ATPase activity of G2ALT requires Mg2+ and Na+ ions, while Zn2+ and Al3+ stimulate the activity. Cd2+ and Ag+ reduced the activity and Li+, Cu2+, and Co2+ inhibited the activity. Known inhibitors of most ATPases, like such as β-mercaptoethanol and ouabain, also inhibited the activity of the G2ALT. These biochemical characterizations suggested that G2ALT belongs to the PP-loop ATPase superfamily and it can be responsible for aluminium tolerance in A. gonensis G2.  相似文献   

20.
The calcium-sensitive forms of adenylyl cyclases (AC) have been revealed in the majority of vertebrate and invertebrate animals, as well as in several representatives of unicellular organisms, including infusoria. We have found for the first time that the AC activity in the infusorian Tetrahymena pyriformis changes in the presence of calcium ions. Calcium ions at concentrations of 0.2–20 μM stimulated the activity of this enzyme, with the maximum of the stimulatory effect being observed at 2 μM Ca2+. At a concentration of 100 μM and higher, the calcium cations inhibited the AC activity. Antagonists of calmodulin W-5 and W-7 at concentrations of 20–100 μM decreased the stimulatory effect of 5 μM Ca2+, while at the higher concentrations inhibited it completely. Another calmodulin antagonist, chloropromazine, decreased the Ca2+-stimulated AC activity only at concentrations of 200–1000 μM. The stimulatory effect of serotonin, EGF, and cAMP on AC activity was enhanced in the presence of 5 μM Ca2+. The stimulatory effect of EGF, cAMP, and insulin on AC was decreased in the presence of 100 μM Ca2+, while the effect of cAMP was also observed in the presence of calmodulin antagonists (500 μM). At the same time, stimulatory effect of D-glucose did not change in the presence of Ca2+ and calmodulin antagonists. The obtained data indicate that, in the infusorian T. pyriformis, there are calcium-sensitive forms of AC that can be stimulated by EGF, cAMP, insulin, and serotonin.  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