首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
Renal hyperosmotic conditions may produce reactive oxygen species, which could have a deleterious effect on the enzymes involved in osmoregulation. Hydrogen peroxide was used to provoke oxidative stress in the environment of betaine aldehyde dehydrogenase in vitro. Enzyme activity was reduced as hydrogen peroxide concentration was increased. Over 50% of the enzyme activity was lost at 100 μM hydrogen peroxide at two temperatures tested. At pH 8.0, under physiological ionic strength conditions, peroxide inhibited the enzyme. Initial velocity assays of betaine aldehyde dehydrogenase in the presence of hydrogen peroxide (0-200 μM) showed noncompetitive inhibition with respect to NAD(+) or to betaine aldehyde at saturating concentrations of the other substrate at pH 7.0 or 8.0. Inhibition data showed that apparent V(max) decreased 40% and 26% under betaine aldehyde and NAD(+) saturating concentrations at pH 8.0, while at pH 7.0 V(max) decreased 40% and 29% at betaine aldehyde and NAD(+) saturating concentrations. There was little change in apparent Km(NAD) at either pH, while Km(BA) increased at pH 7.0. K(i) values at pH 8 and 7 were calculated. Our results suggest that porcine kidney betaine aldehyde dehydrogenase could be inhibited by hydrogen peroxide in vivo, thus compromising the synthesis of glycine betaine.  相似文献   

2.
Cellulose and hemicellulose are the most abundant andmajor constituents of plants and the potential rawmaterials for enzyme hydrolysis related biotechnologicalprocesses. The cellulase and hemicellulase have beenconsidered as two of the most important enzymes inbiotechnology market which can hydrolyze cellulosic ma-terials to fermentable sugars for various biotechnologicalpurposes [1,2], and can also be used in textile- and paper-making industry [3]. It is important to use these enzymesunder re…  相似文献   

3.
Nicotinamide adenine dinucleotide (NAD)-dependent d(minus)-and l(plus)-lactate dehydrogenases have been partially purified 89- and 70-fold simultaneously from cell-free extracts of Pediococcus cerevisiae. Native molecular weights, as estimated from molecular sieve chromatography and electrophoresis in nondenaturing polyacrylamide gels, are 71,000 to 73,000 for d(minus)-lactate dehydrogenase and 136,000 to 139,000 for l(plus)-lactate dehydrogenase. Electrophoresis in sodium dodecyl sulfate-containing gels reveals subunits with approximate molecular weights of 37,000 to 39,000 for both enzymes. By lowering the pyruvate concentration from 5.0 to 0.5 mM, the pH optimum for pyruvate reduction by d(minus)-lactate dehydrogenase decreases from pH 8.0 to 3.6. However, l(plus)-lactate dehydrogenase displays an optimum for pyruvate reduction between pH 4.5 and 6.0 regardless of the pyruvate concentration. The enzymes obey Michaelis-Menten kinetics for both pyruvate and reduced NAD at pH 5.4 and 7.4, with increased affinity for both substrates at the acid pH. alpha-Ketobutyrate can be used as a reducible substrate, whereas oxamate has no inhibitory effect on lactate oxidation by either enzyme. Adenosine triphosphate causes inhibition of both enzymes by competition with reduced NAD. Adenosine diphosphate is also inhibitory under the same conditions, whereas NAD acts as a product inhibitor. These results are discussed with relation to the lactate isomer production during the growth cycle of P. cerevisiae.  相似文献   

4.
1) Glucose dehydrogenase from Bacillus megaterium has been purified to a specific activity of 550 U per mg protein. The homogeneity of the purified enzyme was demonstrated by gel electrophoresis and isoelectric focusing. 2) The amino acid composition has been determined. 3) The molecular weight of the native enzyme was found to be 116000 by gel permeation chromatography, in good agreement with the values of 120000 and 118000, which were ascertained electrophoretically according to the method of Hedrick and Smith and by density gradient centrifugation, respectively. 4) In the presence of 0.1% sodium dodecylsulfate and 8M urea, the enzyme dissociates into subunits with a molecular weight of 30000 as determined by dodecylsulfate gel electrophoresis. These values indicate that the native enzyme is composed of four polypeptide chains, each probably possessing one coenzyme binding site, which can be concluded from fluorescent titration of the NADH binding sites. 5) In polyacrylamide disc electrophoresis, samples of the purified enzyme exhibit three bands of activity, which present the native (tetrameric) form of glucose dehydrogenase and two monomeric forms (molecular weight 30000), arising under the conditions of pH and ionic strength of this method. 6) The enzyme shows a sharp pH optimum at pH 8.0 in Tris/HCl buffer, and a shift of the pH optimum to pH 9.0 in acetate/borate buffer. The limiting Michaelis constant at pH 9.0 for NAD is 4.5 mM and 47.5 mM for glucose. The dissociation constant for NAD is 0.69 mM. 7) D-Glucose dehydrogenase is highly specific for beta-D-glucose and is capable of using either NAD or NADP. The enzyme is insensitive to sulfhydryl group inhibitors, heavy metal ions and chelating agents.  相似文献   

5.
In the presence of NAD(+) the acylation by 1,3-diphosphoglycerate of the four active sites of pig muscle d-glyceraldehyde 3-phosphate dehydrogenase can be monitored at 365nm by the disappearance of the absorption band present in the binary complex of NAD(+) and the enzyme. A non-specific salt effect decreased the acylation rate 25-fold when the ionic strength was increased from 0.10 to 1.0. This caused acylation to be the rate-limiting process in the enzyme-catalysed reductive dephosphorylation of 1,3-diphosphoglycerate at high ionic strength at pH8. The salt effect permitted investigation of the acylation over a wide range of conditions. Variation of pH from 5.4 to 8.6 produced at most a two-fold change in the acylation rate. One proton was taken up per site acylated at pH8.0. By using a chromophoric H(+) indicator the rate of proton uptake could be monitored during the acylation and was also almost invariant in the pH range 5.5-8.5. Transient kinetic studies of the overall enzyme-catalysed reaction indicated that acylation was the process involving proton uptake at pH8.0. The enzyme mechanism is discussed in the light of these results.  相似文献   

6.
1. The kinetics of oxidation of l-glycerol 3-phosphate by NAD(+) and of reduction of dihydroxyacetone phosphate by NADH catalysed by rabbit muscle glycerol 3-phosphate dehydrogenase were studied over the range pH6-9. 2. The enzyme was found to catalyse the oxidation of glyoxylate by NAD(+) at pH8.0 and the kinetics of this reaction were also studied. 3. The results are consistent with a compulsory mechanism of catalysis for glycerol 3-phosphate oxidation and dihydroxyacetone phosphate reduction in the intermediate regions of pH, but modifications to the basic mechanism are required to fully explain results at the extremes of the pH range, with these substrates and for glyoxylate oxidation at pH8.0.  相似文献   

7.
NAD glycohydrolase, or NADase (NAD+ glycohydrolase, EC 3.2.2.5) was solubilized with porcine pancreatic lipase from isolated fractions of microsomes and plasma membranes obtained from rat livers. The enzyme from each organelle was further purified by DEAE-cellulose chromatography, gel filtration and isoelectric focusing. The solubilized, partially purified enzymes had similar molecular weights, pH-activity profiles and Km values. Marked charge heterogeneity was observed for the microsomal enzyme on isoelectric focusing between pH 6 and 8 with maximum activity focusing at pH 8.0. Plasma membrane NADase displayed a single peak at pH 6.7. Treatment of the partially purified microsomal or plasma membrane enzyme with neuraminidase resulted in a single peak of activity on isoelectric focusing (pH 3.5--10) with a pI of 9.2. Polyacrylamide gel electrophoresis of either NADase revealed a periodate-Schiff positive band which was coincident with enzyme activity. Compositional analyses of the microsomal enzyme focusing at pH 8.0 confirmed the presence of hexoses, hexosamines and sialic acid. Differences in carbohydrate composition might be important in determining the subcellular distribution of this enzyme.  相似文献   

8.
Thermal unfolding parameters were determined for a two-domain tetrameric enzyme, phosphorylating D-glyceraldehyde-3-phosphate dehydrogenase (GAPDH), and for its isolated NAD(+)-binding domain. At pH 8.0, the transition temperatures (t(max)) for the apoforms of the native Bacillus stearothermophilus GAPDH and the isolated domain were 78.3 degrees C and 61.9 degrees C, with calorimetric enthalpies (DeltaH(cal)) of 4415 and 437 kJ/mol (or 30.7 and 22.1 J/g), respectively. In the presence of nearly saturating NAD(+) concentrations, the t(max) and the DeltaH(cal) increased by 13.6 degrees C and by 2365 kJ/mol, respectively, for the native apoenzyme, and by 2.8 degrees C and 109 kJ/mol for the isolated domain. These results indicate that interdomain interactions are essential for NAD(+) to produce its stabilizing effect on the structure of the native enzyme. The thermal stability of the isolated NAD(+)-binding domain increased considerably upon transition from pH 6.0 to 8.0. By contrast, native GAPDH exhibited greater stability at pH 6.0; similar pH-dependencies of thermal stability were displayed by GAPDHs isolated from rabbit muscle and Escherichia coli. The binding of NAD(+) to rabbit muscle apoenzyme increased t(max) and DeltaH(cal) and diminished the widths of the DSC curves; the effect was found to grow progressively with increasing coenzyme concentrations. Alkylation of the essential Cys149 with iodoacetamide destabilized the apoenzyme and altered the effect of NAD(+). Replacement of Cys149 by Ser or by Ala in the B. stearothermophilus GAPDH produced some stabilization, the effect of added NAD(+) being basically similar to that observed with the wild-type enzyme. These data indicate that neither the ion pairing between Cys149 and His176 nor the charge transfer interaction between Cys149 and NAD(+) make any significant contribution to the stabilization of the enzyme's native tertiary structure and the accomplishment of NAD(+)-induced conformational changes. The H176N mutant exhibited dramatically lower heat stability, as reflected in the values of both DeltaH(cal) and t(max). Interestingly, NAD(+) binding resulted in much wider heat capacity curves, suggesting diminished cooperativity of the unfolding transition.  相似文献   

9.
Some properties of ADP-ribose transferase, and its reaction product, from BHK-21/C13 cells are described. Enzyme activity was found almost exclusively in nuclei (90%), with the remaining 10% located in the cytosolic fraction. The nuclear enzyme is chromatin-bound and requires bivalent cations, preferably Mg2+, a pH of 8.0 and a temperature of 25 degrees C for optimal activity. Chromatin preparations incorporated radioactivity from [14C]NAD+ into acid-insoluble material for about 60 min. Kinetics for substrate NAD+ utilization were not of Michaelis--Menten type; biphasic kinetics were shown from a double-reciprocal plot (1/reaction velocity against 1/[NAD+]) and from a 'Hofstee' plot (reaction velocity/[NAD+] against reaction velocity). The transferase is unstable in the absence of Mg2+ ions. It is inhibited by thymidine, nicotinamide and nicotinamide analogues, but not by ATP, which stimulates it at concentrations of 5 mM and above. The enzyme requires thiol groups for activity; it is readily inhibited by N-ethylmaleimide at 0.5 mM. The product of the reaction is stable under acid conditions at temperatures up to 25 degrees C, but it is hydrolysed by HClO4 at 70 degrees C. It is resistant to NaOH, but is cleaved from its attachment to protein with alkali into trichloroacetic acid-insoluble and -soluble components. On the basis of Cs2SO4- density-gradient analysis under denaturing conditions (gradients included urea and guanidinium hydrochloride), and analysis of the reaction product directly on hydroxyapatite, we conclude that most of the radioactive ADP-ribose residues are firmly bound to protein, presumably in covalent linkage. Hydroxyapatite-chromatographic analysis of ADP-ribose residues released from protein by alkaline digestion showed a spectrum of molecular sizes including mono-, oligo- and poly-(ADP-ribose), when chromatin was incubated initially with [14C]NAD+ for 10 min and then for a further 30 min after addition of excess non-radioactive NAD+, only about 10% of the radioactive mono-(ADP-ribose) could be 'chased' into longer-chain molecules. Hydroxyapatite analysis was also used to show that, whereas all ADP-ribose residues were released from protein with NaOH, only 50% of them were susceptible to hydroxylamine. These hydroxylamine-sensitive residues included all size classes, although mono-(ADP-ribose) predominated. Finally, there was an approximately equal distribution of ADP-ribose incorporated into HCl-soluble proteins (including the histones) and HCl-insoluble proteins (including the non-histone proteins) when chromatin was incubated with NAD+ up to 0.5 mM, but at higher NAD+ concentrations more ADP-ribose was incorporated into the HCl-soluble fraction (82% at 4.0 mM-NAD+).  相似文献   

10.
UDP-galactose 4-epimerase from Escherichia coli is a homodimer of 39 kDa subunit with non-covalently bound NAD acting as cofactor. The enzyme can be reversibly reactivated after denaturation and dissociation using 8 M urea at pH 7.0. There is a strong affinity between the cofactor and the refolded molecule as no extraneous NAD is required for its reactivation. Results from equilibrium denaturation using parameters like catalytic activity, circular-dichroism, fluorescence emission (both intrinsic and with extraneous fluorophore 1-aniline 8-naphthalene sulphonic acid), 'reductive inhibition' (associated with orientation of NAD on the native enzyme surface), elution profile from size-exclusion HPLC and light scattering have been compiled here. These show that inactivation, integrity of secondary, tertiary and quaternary structures have different transition mid-points suggestive of non-cooperative transition. The unfolding process may be broadly resolved into three parts: an active dimeric holoenzyme with 50% of its original secondary structure at 2.5 M urea; an active monomeric holoenzyme at 3 M urea with only 40% of secondary structure and finally further denaturation by 6 M urea leads to an inactive equilibrium unfolded state with only 20% of residual secondary structure. Thermodynamical parameters associated with some transitions have been quantitated. The results have been discussed with the X-ray crystallographic structure of the enzyme.  相似文献   

11.
GDP-Mannose dehydrogenase catalyzes the formation of GDP-mannuronic acid, which is the monomeric unit from which the polysaccharide alginate is formed. Alginate is secreted by the pathogenic bacterium Pseudomonas aeruginosa and is believed to play an important role in the bacteria's resistance to antibiotics and the host immune response. We have characterized the kinetic behavior of GDP-mannose dehydrogenase in detail. The enzyme displays cooperative behavior with respect to NAD(+) binding, and phosphate and GMP act as allosteric effectors. Binding of the allosteric effectors causes the Hill coefficient for NAD(+) binding to decrease from 6 to 1, decreases K(1/2) for NAD(+) by a factor of 10, and decreases V(max) by a factor of 2. The cooperative binding of NAD(+) is also sensitive to pH; deprotonation of two residues with identical pK's of 8.0 is required for maximally cooperative behavior. The kinetic behavior of GDP-mannose dehydrogenase suggests that it must be at least hexameric under turnover conditions; however, dynamic light-scattering measurements do not provide a clear determination of the size of the active enzyme complex.  相似文献   

12.
M R Eftink  K Bystr?m 《Biochemistry》1986,25(21):6624-6630
The association of the coenzyme NAD+ to liver alcohol dehydrogenase (LADH) is known to be pH dependent, with the binding being linked to the shift in the pK of some group on the protein from a value of 9-10, in the free enzyme, to 7.5-8 in the LADH-NAD+ binary complex. We have further characterized the nature of this linkage between NAD+ binding and proton dissociation by studying the pH dependence (pH range 6-10) of the proton release, delta n, and enthalpy change, delta Ho(app), for formation of both binary (LADH-NAD+) and ternary (LADH-NAD+-I, where I is pyrazole or trifluoroethanol) complexes. The pH dependence of both delta n and delta Ho(app) is found to be consistent with linkage to a single acid dissociating group, whose pK is perturbed from 9.5 to 8.0 upon NAD+ binding and is further perturbed to approximately 6.0 upon ternary complex formation. The apparent enthalpy change for NAD+ binding is endothermic between pH 7 and pH 10, with a maximum at pH 8.5-9.0. The pH dependence of the delta Ho(app) for both binary and ternary complex formation is consistent with a heat of protonation of -7.5 kcal/mol for the coupled acid dissociating group. The intrinsic enthalpy changes for NAD+ binding and NAD+ plus pyrazole binding to LADH are determined to be approximately 0 and -11.0 kcal/mol, respectively. Enthalpy change data are also presented for the binding of the NAD+ analogues adenosine 5'-diphosphoribose and 3-acetylpyridine adenine dinucleotide.  相似文献   

13.
By use of a simple device for continuous CO2 gassing of Ureaplasma urealyticum cultures growing in a liquid medium, we have been able to separate some of the effects of urea, CO2, ammonia, and pH on growth. The CO2 acted as a superior buffer in the pH range 5.7 to 6.8, which is optimal for Ureaplasma growth. It was, therefore, possible to observe the effect of repeated additions of urea to the culture without alkalinization of the growth medium. We found that the repeated additions of urea did not enhance Ureaplasma growth, and the resultant accumulation of ammonium ions (greater than 2,000 microng/ml) did not cause more rapid death under these conditions. By abruptly changing the gaseous environment from CO2 to N2, it was possible to cause a rapid pH change in the culture to a value above 8.0. This resulted in a more rapid death of the organisms.  相似文献   

14.
Y Chi  T K Kumar  H M Wang  M C Ho  I M Chiu  C Yu 《Biochemistry》2001,40(25):7746-7753
The thermodynamic parameters characterizing the conformational stability of the human acidic fibroblast growth factor (hFGF-1) have been determined by isothermal urea denaturation and thermal denaturation at fixed concentrations of urea using fluorescence and far-UV CD circular dichroism (CD) spectroscopy. The equilibrium unfolding transitions at pH 7.0 are adequately described by a two-state (native <--> unfolded state) mechanism. The stability of the protein is pH-dependent, and the protein unfolds completely below pH 3.0 (at 25 degrees C). hFGF-1 is shown to undergo a two-state transition only in a narrow pH range (pH 7.0-8.0). Under acidic (pH <6.0) and basic (pH >8.0) conditions, hFGF-1 is found to unfold noncooperatively, involving the accumulation of intermediates. The average temperature of maximum stability is determined to be 295.2 K. The heat capacity change (DeltaC(p)()) for the unfolding of hFGF-1 is estimated to be 2.1 +/- 0.5 kcal.mol(-1).K(-1). Temperature denaturation experiments in the absence and presence of urea show that hFGF-1 has a tendency to undergo cold denaturation. Two-dimensional (1)H-(15)N HSQC spectra of hFGF-1 acquired at subzero temperatures clearly show that hFGF-1 unfolds under low-temperature conditions. The significance of the noncooperative unfolding under acidic conditions and the cold denaturation process observed in hFGF-1 are discussed in detail.  相似文献   

15.
Pasteurella multocida toxin is a potent mitogen that is believed to act intracellularly. On transverse urea gradient gels at pH 8.0 the toxin displayed one major unfolding transition at 4 M urea. However, at pH 6.1 the unfolding transition took place at 3.5 M urea. Circular dichroism spectra also indicated that a structural change took place at acidic pH. In addition it was found that the toxin that had been denatured in 8 M urea refolded in solution with a high recovery of biological activity. These findings are discussed in terms of the likely domain structure of the P. multocida toxin.  相似文献   

16.
The NAD kinase gene (nadK) of Sphingomonas sp. A1 was cloned and then overexpressed in Escherichia coli, and the gene product (NadK) was purified from the E. coli cells through five steps with a 25% yield of activity. NadK was a homodimer of 32 kDa subunits, utilized ATP or other nucleoside triphosphates, but not inorganic polyphosphates, as phosphoryl donors for the phosphorylation of NAD, most efficiently at pH 8.0 and 50-55 degrees C, and was designated as ATP-NAD kinase (NadK). NadK showed no NADH kinase activity and was slightly inhibited by NADP(H). Precursors for NAD biosynthesis such as quinolinic acid, nicotinic acid mononucleotide, nicotinic acid adenine dinucleotide, and nicotinic acid had no effect on the NadK activity, as observed in the cases of the NAD kinases of Micrococcus flavus, Mycobacterium tuberculosis, and E. coli. Taken together with the report that the NAD kinase of Bacillus subtilis is activated by quinolinic acid [J. Bacteriol. 185 (2003) 4844], it is indicated that the regulatory patterns of NAD kinases differ even among bacterial NAD kinases.  相似文献   

17.
The reactions of the NAD radical (NAD.) with ferric horseradish peroxidase and with compounds I and II were investigated by pulse radiolysis. NAD. reacted with the ferric enzyme and with compound I to form the ferrous enzyme and compound II with second-order rate constants of 8 X 10(8) and 1.5 X 10(8) M-1 s-1, respectively, at pH 7.0. In contrast, no reaction of NAD. with native compound II at pH 10.0 nor with diacetyldeutero-compound II at pH 5.0-8.0 could be detected. Other reducing species generated by pulse radiolysis, such as hydrated electron (eaq-), superoxide anion (O2-), and benzoate anion radical, could not reduce compound II of the enzyme to the ferric state, although the methylviologen radical reduced it. The results are discussed in relation to the mechanism of catalysis of the one-electron oxidation of substrates by peroxidase.  相似文献   

18.
Phosphite dehydrogenase (PTDH) catalyzes the NAD-dependent oxidation of phosphite to phosphate, a reaction that is 15 kcal/mol exergonic. The enzyme belongs to the family of D-hydroxy acid dehydrogenases. Five other family members that were analyzed do not catalyze the oxidation of phosphite, ruling out the possibility that this is a ubiquitous activity of these proteins. PTDH does not accept any alternative substrates such as thiophosphite, hydrated aldehydes, and methylphosphinate, and potential small nucleophiles such as hydroxylamine, fluoride, methanol, and trifluoromethanol do not compete with water in the displacement of the hydride from phosphite. The pH dependence of k(cat)/K(m,phosphite) is bell-shaped with a pK(a) of 6.8 for the acidic limb and a pK(a) of 7.8 for the basic limb. The pK(a) of 6.8 is assigned to the second deprotonation of phosphite. However, whether the dianionic form of phosphite is the true substrate is not clear since a reverse protonation mechanism is also consistent with the available data. Unlike k(cat)/K(m,phosphite), k(cat) and k(cat)/K(m,NAD) are pH-independent. Sulfite is a strong inhibitor of PTDH that is competitive with respect to phosphite and uncompetitive with respect to NAD(+). Incubation of the enzyme with NAD(+) and low concentrations of sulfite results in a covalent adduct between NAD(+) and sulfite in the active site of the enzyme that binds very tightly. Fluorescent titration studies provided the apparent dissociation constants for NAD(+), NADH, sulfite, and the sulfite-NAD(+) adduct. Substrate isotope effect studies with deuterium-labeled phosphite resulted in small normal isotope effects (1.4-2.1) on both k(cat) and k(cat)/K(m,phosphite) at pH 7.25 and 8.0. Solvent isotope effects (SIEs) on k(cat) are similar in size; however, the SIE of k(cat)/K(m,phosphite) at pH 7.25 is significantly larger (4.4), whereas at pH 8.0, it is the inverse (0.6). The pH-rate profile of k(cat)/K(m,phosphite), which predicts that the observed SIEs will have a significant thermodynamic origin, can account for these effects.  相似文献   

19.
Proteins in extracts from cotyledons, hypocotyls, and roots of 5-d-old, dark-grown soybean (Glycine max L. Merr. cv Williams) seedlings were separated by polyacrylamide gel electrophoresis. Three isoforms of glutamate dehydrogenase (GDH) were resolved and visualized in gels stained for GDH activity. Two isoforms with high electrophoretic mobility, GDH1 and GDH2, were in protein extracts from cotyledons and a third isoform with the lowest electrophoretic mobility, GDH3, was identified in protein extracts from root and hypocotyls. Subcellular fractionation of dark-grown soybean tissues demonstrated that GDH3 was associated with intact mitochondria. GDH3 was purified to homogeneity, as determined by native and sodium dodecyl sulfate-polyacrylamide gels. The isoenzyme was composed of a single 42-kD subunit. The pH optima for the reductive amination and the oxidative deamination reactions were 8.0 and 9.3, respectively. At any given pH, GDH activity was 12- to 50-fold higher in the direction of reductive amination than in the direction of the oxidative deamination reaction. GDH3 had a cofactor preference for NAD(H) over NADP(H). The apparent Michaelis constant values for [alpha]-ketoglutarate, ammonium, and NADH at pH 8.0 were 3.6, 35.5, and 0.07 mM, respectively. The apparent Michaelis constant values for glutamate and NAD were 15.8 and 0.10 mM at pH 9.3, respectively. To our knowledge, this is the first biochemical and physical characterization of a purified mitochondrial NAD(H)-dependent GDH isoenzyme from soybean.  相似文献   

20.
A new type of glucose dehydrogenase was purified from vegetative cells of Bacillus megaterium IAM1030. The characteristics of the vegetative-cell enzyme were investigated and compared with the enzyme from sporulating cells of B. megaterium IWG3. They are very similar in the following points: molecular size (Mr 120,000), subunit composition (homo tetramer), pH-activity profile with an optimum pH at around 8, pH-stability profile with a stable pH range of 6.0–7.5 (at 25°C, for 30 min), substrate specificity (specific for d-glucose and 2-deoxy-d-glucose), and the affinity for glucose (a Km value of 11–12 mM at pH 8.0, 2.5 mM NAD). They are a little different in the following points: slower mobility for the vegetative-cell enzyme in polyacrylamide-gel electrophoresis at pH 8, immunological determinants (some of them are common), and higher heat resistance for the vegetative-cell enzyme at pH 6.5. They are quite different in their affinity for NAD and NADP. The Km values for NAD are 0.1 mM for the vegetative-cell enzyme and 1.0 mM for the spore enzyme, while the values for NADP are 7.1 mM for the vegetative-cell enzyme and 0.09 mM for the spore enzyme, at pH 8.0, 0.1 M d-glucose. These results suggest that B. megaterium has at least two types of glucose dehydrogenase.  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