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 62 毫秒
1.
Phosphoglyceromutase (PGM) from chicken breast muscle was titrated with p-mercuribenzoate (PMB), 5,5'-dithiobisnitrobenzoate (Nbs2), N-ethylmaleimide (NEM), iodoacetate and iodoacetamide. The effect of all of the sulfhydryl reagents, with the exception of NEM was to cause a loss in enzymatic activity. Addition of KCN following reaction with Nbs2 resulted in the recovery of a small amount of enzymatic activity. In the absence of substrate (3-phosphoglyceric acid) or cofactor (2,3-diphosphoglyceric acid) and in the presence or absence of 6 M guanidine hydrochloride, six sulfhydryl groups per mole of enzyme were titrated with PMB.  相似文献   

2.
Bacillus stearothermophilus phosphoglycerate mutase (PGM), which interconverts 2- and 3-phosphoglyceric acid (PGA), does not require 2,3-diphosphoglyceric acid for activity. However, this enzyme does have an absolute and specific requirement for Mn(2+) ions for catalysis. Here we report the crystal structure of this enzyme complexed with 3PGA and manganese ions to 1.9 A resolution; this is the first crystal structure of a diphosphoglycerate-independent PGM to be determined. This information, plus the location of the two bound Mn(2+) ions and the 3PGA have allowed formulation of a possible catalytic mechanism for this PGM. In this mechanism Mn(2+) ions facilitate the transfer of the substrate's phosphate group to Ser62 to form a phosphoserine intermediate. In the subsequent phosphotransferase part of the reaction, the phosphate group is transferred from Ser62 to the O2 or O3 positions of the reoriented glycerate to yield the PGA product. Site-directed mutagenesis studies were used to confirm our mechanism and the involvement of specific enzyme residues in Mn(2+) binding and catalysis.  相似文献   

3.
Phosphoglycerate phosphomutase has been purified to homogeneity from vegetative cells and germinated spores of Bacillus megaterium, and the spore and cell enzymes appear identical. The enzyme is a monomer of molecular weight 61,000. The compound 2,3-diphosphoglyceric acid is not required for activity, but the enzyme has an absolute and specific requirement for Mn2+. The enzyme is inhibited by ethylenediaminetetraacetate and sulfhydryl reagents, has a pH optimum of about 8.0, and has Km values for 3-phosphoglyceric acid and Mn2+ of 5 x 10(-4) and 4 x 10(-5) M, respectively.  相似文献   

4.
Effects of the substrate and the coenzyme on the crystalline yeast phosphoglyceric acid mutase activity have been investigated. Lineweaver-Burk plots at different concentrations of the substrate (d-3-phosphoglyceric acid: 3×10?7 to 8×10?3m) and the coenzyme (d-2, 3-diphosphoglyceric acid: 8×10?7 to 10?5m) change in such a way to indicate the involvement of an enzyme-substrate-coenzyme ternary complex as an active intermediate in the enzymic reaction process. It is concluded that the reaction catalyzed by the yeast enzyme follows the sequential pathway and that a phosphorylated enzyme does not participate as an obligatory intermediate in the reaction mechanism, if it occurs. Kinetic studies indicate Km values of 6×10?4m for d-3-phosphoglyceric acid and 8×10?7m for d-2, 3-diphosphoglyceric acid. The substrate is a competitive inhibitor of the coenzyme with a Ksi (inhibition constant) of 3.2×10?3m. The coenzyme inhibition is not observed at concentration tested. A kinetic treatment to determine the mechanism of the enzyme reaction from the experimental data which are obtaind in the range of inhibitory substrate concentrations is presented.  相似文献   

5.
A simple two-step procedure for purification of enolase from germinated spores or vegetative cells of Bacillus megaterium is described. The procedure resulted in a 1,200-fold purification with production of homogeneous enzyme in approximately 75% yield; the enzymes from spores and cells seemed identical. The molecular weight of the native enzyme was 335,000, with a subunit molecular weight of 42,000. The enzyme required Mg2+ and was inhibited by ethylenediaminetetraacetic acid and fluoride ions. The Michaelis constants for 2-phosphoglyceric acid and Mg2+ were 7.1 X 10(-4) and 4.7 X 10(-4) M, respectively.  相似文献   

6.
S Liu  M J Gresser  A S Tracey 《Biochemistry》1992,31(10):2677-2685
The formation of complexes of vanadate with 2-phosphoglycerate and 3-phosphoglycerate have been studied using 51V nuclear magnetic resonance spectroscopy. Signals attributed to two 2,3-diphosphoglycerate analogues, 2-vanadio-3-phosphoglycerate and 2-phospho-3-vanadioglycerate, were detected but were not fully resolved from signals of inorganic vanadate and the anhydride formed between vanadate and the phosphate ester moieties of the individual phosphoglycerates. Equilibrium constants for formation of the two 2,3-bisphosphate analogues were estimated as 2.5 M-1 for 2-vanadio-3-phosphoglycerate and 0.2 M-1 for 2-phospho-3-vanadioglycerate. The results of the binding study are fully consistent with non-cooperativity in the binding of vanadiophosphoglycerate to the two active sites of phosphoglycerate mutase (PGM). 2-Vanadio-3-phosphoglycerate was found to bind to the dephospho form of phosphoglycerate mutase with a dissociation constant of about 1 x 10(-11) M at pH 7 and 7 x 10(-11) M at pH 8. Three signals attributed to histidine residues were observed in the 1H NMR spectrum of phosphoglycerate mutase. Two of these signals and also an additional signal, tentatively attributed to a tryptophan, underwent a chemical shift change when the vanadiophosphoglycerate complex was bound to the enzyme. The results obtained here are in accord with these vanadate-phosphoglycerate complexes being much more potent inhibitors of phosphoglycerate mutase than either monomeric or dimeric vanadate. The dissociation constant of 10(-11) M for 2-vanadio-3-phosphoglycerate is about 4 orders of magnitude smaller than the Km for PGM, a result in accordance with the vanadiophosphoglycerates being transition state analogues for the phosphorylation of PGM by 2,3-diphosphoglycerate.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

7.
A neuraminidase activity in myelin isolated from adult rat brains was examined. The enzyme activity in myelin was first compared with that in microsomes using N-acetylneuramin(alpha 2----3)lactitol (NL) as a substrate. In contrast to the microsomal neuraminidase which exhibited a sharp pH dependency for its activity, the myelin enzyme gave a very shallow pH activity curve over a range between 3.6 and 5.9. The myelin enzyme was more stable to heat denaturation (65 degrees C) than the microsomal enzyme. Inhibition studies with a competitive inhibitor, 2,3-dehydro-2-deoxy-N-acetylneuraminic acid, showed the Ki value for the myelin neuraminidase to be about one-fifth of that for the microsomal enzyme (1.3 X 10(-6) M versus 6.3 X 10(-6) M). The apparent Km values for the myelin and the microsomal enzyme were 1.3 X 10(-4) M and 4.3 X 10(-4) M, respectively. An enzyme preparation that was practically devoid of myelin lipids was then prepared and its substrate specificity examined. The "delipidated enzyme" could hydrolyze fetuin, NL, and ganglioside substrates, including GM1 and GM2. When the delipidated enzyme was exposed to high temperature (55 degrees C) or low pH (pH 2.54), the neuraminidase activities toward NL and GM3 decreased at nearly the same rate. Both fetuin and 2,3-dehydro-2-deoxy-N-acetylneuraminic acid inhibited NL and GM3 hydrolysis. With 2,3-dehydro-2-deoxy-N-acetylneuraminic acid, inhibition of NL was greater than that of GM3; however, the Ki values for each substrate were almost identical. GM3 and GM1 also competitively inhibited the hydrolysis of NL and NL similarly inhibited GM3 hydrolysis by the enzyme. These results indicate that rat brain myelin has intrinsic neuraminidase activities toward nonganglioside as well as ganglioside substrates, and that these two enzyme activities are likely catalyzed by a single enzyme entity.  相似文献   

8.
The neuraminidase associated with the bifunctional protein, hemagglutinin-neuraminidase, of influenza virus has been characterized. The enzyme has a pH optimum of 4.5, does not require Ca2+ and is inactivated (98%) by incubation at 50 degrees C. The enzyme has a Km of 2.00 X 10(-3) M and 0.06 X 10(-3) M with the substrates 2-(3-methoxyphenyl)-N-acetylneuraminic acid and fetuin, respectively. The Ki is 400 X 10(-6) with the inhibitor 2-deoxy-2,3-dehydro-N-acetylneuraminic acid. The incorporation of labeled cysteine, valine and leucine in the hemagglutinin-neuraminidase protein is different from that of viral neuraminidase. A comparison of the properties of the neuraminidase associated with protein hemagglutinin-neuraminidase with that of viral neuraminidase or sialidase showed that the former is biochemically different and an antigenically distinct enzyme. The unique feature of the new enzyme is that it has the hemagglutinin activity as well. The two biological activities could not be separated from each other in all systems used. Apparently, protein hemagglutinin-neuraminidase is genetically transferable and it is detectable in a laboratory recombinant virus E-2971 (H3 Aichi X N7). These results suggest that protein hemagglutinin-neuraminidase is a unique surface protein of the influenza virus A/Aichi/2/68 (H3N2).  相似文献   

9.
N-Bromoacetylethanolamine phosphate rapidly and irreversibly inactivates rabbit muscle phosphoglycerate mutase. At high molar ratios of reagent to enzyme, loss of activity (both mutase and phosphatase) approximates pseudo-first order kinetics. A rate-saturation effect is observed with half-maximal rate of inactivation occurring at 0.32 mM reagent, a value close to the Km for 3-phosphoglyceric acid. This datum and the dissociation constant of the 2,3-bisphosphoglycerate-enzyme complex, as determined from inactivation kinetics in the presence of the bisphosphate, suggest that the reagent reacts at the substrate binding site. Inactivation results from the covalent incorporation of about 0.8 mol of reagent/mol of catalytic subunit as determined with 14C-labeled reagent. Incorporation is negligible in the presence of substrate and is reduced 8-fold in the presence of 6 M urea. From amino acid analyses on acid hydrolysates of the inactivated enzyme, we have identified a sulfhydryl group as the site of alkylation. A peptide containing the essential sulfhydryl group has been isolated from a tryptic digest of the enzyme inactivated with labeled reagent; its amino acid composition is Trp1, Lys1,-Cys(Cm)1, Asp1, Ser1, Glu2, Gly1, Ala1, Leu1, Phe2.  相似文献   

10.
1. The three isozymes of glycerate-2,3-P2 dependent phosphoglycerate mutase present in tissues of mammals and reptiles were inactivated by both treatment with diethylpyrocarbonate and photooxidation with rose bengal. 2. Inactivation of type M isozyme purified from rabbit muscle was complete when two histidine residues per enzyme subunit were carboethoxylated. Hydroxylamine removed the carboethoxy groups, with partial recovery of the enzymatic activity. The cofactor protected the enzyme against inactivation. 3. The inactivation of rabbit muscle phosphoglycerate mutase by photooxidation with methylene blue and rose bengal was sharply pH dependent. The pH profile of enzyme inactivation followed the titration curve of histidine, suggesting that this amino acid was critical for enzyme activity. Glycerate-2,3-P2 did not protect phosphoglycerate mutase against photoinactivation.  相似文献   

11.
1. The three phosphoglycerate mutase isozymes from mammals (types M, B and MB isozymes) differ in their sensitivity to the - SH group reagents. 2. Rabbit muscle phosphoglycerate mutase (type M isozyme) is reversibly inactivated by tetrathionate, rho-chloromercuribenzoate and Hg2+. 3. Titration with rho-chloromercuribenzoate shows the existence of two sulfhydryl groups per enzyme subunit, the modification of which produces a progressive decline in enzyme activity. 4. The apparent Km values for substrate and cofactor are not affected by tetrathionate treatment. 5. Phosphoglycerate mutase inactivated by tetrathionate and by rho-chloromercuribenzoate is unable to form the functionally active phosphorylenzyme when mixed with glycerate-2,3-P2, and is not protected by the cofactor against heating. 6. Glycerate-2,3-P2 protects against tetrathionate treatment, but fails to protect against Hg2+ and rho-chloromercuribenzoate inactivation.  相似文献   

12.
We have examined the effects of folate compounds and the folate analog amethopterin (methotrexate) as inhibitors of mammalian xanthine oxidase and have found that they offer potent inhibition of the enzyme. We have compared the inhibitory potency of folic acid and its coenzyme derivative tetrahydrofolic acid to that of allopurinol, a known inhibitor of xanthine oxidase, and have demonstrated that folic acid and tetrahydrofolic acid are severalfold more potent than allopurinol as inhibitors of xanthine oxidase. Comparative inhibition constants calculated were 5.0 X 10(-7) M for folic acid. 1.25 X 10(-6) M for tetrahydrofolic acid, and 4.88 X 10(-6) M for allopurinol. Incubation of xanthine oxidase with folic acid at a concentration of 10(-6) M abolished 94% of the enzymic activity within 1 min of incubation with the enzyme. At the same concentration, allopurinol was almost ineffective as an inhibitor of xanthine oxidase. The substrate xanthine protected the enzyme against total inhibition by folic acid. Reversibility of the enzymic inhibition by folic acid was demonstrated. Folic acid-inactivated enzyme was totally regenerated either by filtration through Sephadex G-200 or by precipitation with ammonium sulfate. 2-Amino-4-hydroxypteridine was a poor substrate for the enzyme but a potent inhibitor for the oxidation of xanthine by the enzyme. The inhibition constant calculated was 1.50 X 10(-6) M. In the presence of an excess of xanthine oxidase, neither folic acid nor tetrahydrofolic acid and allopurinol exhibited any change in intensity of their absorbance or in the wavelength of their maximal absorbance that might have been suggestive of substrate utility. The folate analog amethopterin was also determined a potent inhibitor of mammalian xanthine oxidase. The inhibition constant calculated was 3.0 X 10(-5) M.  相似文献   

13.
Cloning and sequencing of a murine cDNA with the entire coding region of 2,3-bisphosphoglycerate mutase is reported, as a prerequisite for further expression studies of this erythroid specific enzyme in Friend mouse erythroleukemia cells. A comparison between species of the deduced amino acid sequences of these proteins shows 20 substitutions between mouse and human and 21 between mouse and rabbit: none of these substitutions are in positions assumed to be in the active site. Amino acid alignment with the other related enzymes, the phosphoglycerate mutases, in combination with crystallographic data from yeast phosphoglycerate mutase, gives some insight into the structure/function correlation for this protein family. Amino acid residues which are most likely critical for either 2,3-bisphosphoglycerate mutase or phosphoglycerate mutase function are pointed out. Concerning the phylogenetic analysis, phosphoglycerate mutases B and M from mammalians appear to have diverged with the yeast enzyme from a common ancestor, before the emergence of the 2,3-bisphosphoglycerate mutases.  相似文献   

14.
The bisphosphatase domain of the rat liver 6-phosphofructo-2-kinase/fructose-2,6-bisphosphatase has been shown to exhibit a structural similarity to yeast phosphoglycerate mutase and human red blood cell 2,3-bisphosphoglycerate mutase including very similar active site sequences with a histidyl residue being involved in phospho group transfer. The liver bifunctional enzyme was found to catalyze the hydrolysis of glycerate 1,3-bisphosphate to glycerate 3-phosphate and inorganic phosphate. The Km for glycerate 1,3-bisphosphate was 320 microM and the Vmax was 11.5 milliunits/mg. Incubation of the rat liver enzyme with [1-32P]glycerate 1,3-bisphosphate resulted in the formation of a phosphoenzyme intermediate, and the labeled amino acid was identified as 3-phosphohistidine. Tryptic and endoproteinase Lys-C peptide maps of the 32P-phosphoenzyme labeled either with [2-32P]fructose 2,6-bisphosphate or [1-32P]glycerate 1,3-bisphosphate revealed that 32P-radioactivity was found in the same peptide, proving that the same histidyl group accepts phosphate from both substrates. Fructose 2,6-bisphosphate inhibited competitively the formation of phosphoenzyme from [1-32P]glycerate 1,3-bisphosphate. Effectors of fructose-2,6-bisphosphatase also inhibited phosphoenzyme formation. Substrates and products of phosphoglycerate mutase and 2,3-bisphosphoglycerate mutase also modulated the activities of the bifunctional enzyme. These results demonstrate that, in addition to a structural homology, the bisphosphatase domain of the bifunctional enzyme has a functional similarity to phosphoglycerate mutase and 2,3-bisphosphoglycerate mutase and support the concept of an evolutionary relationship between the three enzyme activities.  相似文献   

15.
Homogeneous phosphoglycerate kinase from bovine liver possesses a maximum ultraviolet absorption at 278 nm (A 1%,1Cm 280 equals 6.7; Amax/Amin equals 2.26; e280 equals 31.5 mM(-1) X cm(-1). The enzyme consists of about 420 amino-acid residues and is a slightly acidic protein with an isoelectric point of 6.5 as expected from amino-acid analysis. The most notable features of the chemical composition are two tryptophan, 12 methionine and four half-cystine residues per enzyme molecule. Although phosphoglycerate kinases from mammalian tissues are partially similar to each other, clear differences in serine, glutamic acid, glycine, cysteine, valine, leucine, tyrosine, tryptophan and arginine contents were found. Fingerprinting and column chromatography of tryptic digests of the S-carboxymethylated protein confirm the data of amino-acid analysis. Liver phosphoglycerate kinase is inactivated when modified with either p-chloromercuribenzoate or 5,5'dithio-bis(2-nitrobenzoic acid) (Nbs2). The enzyme has two thiol groups available for reaction with Nbs2 under denaturing conditions, one of which is essential for catalysis. After reduction by NaBH4 four cysteine residues per molecule were determined with Nbs2, sugessting the presence of a disulfide bridge. Using sedimentation equilibrium studies, the molecular weight was found to be 49600. Gel filtration yielded values of 43000-50000. By analytical dodecylsulfate-polyacrylamide gel electrophoresis a molecular weight of 45600 was estimated. Inconsistent with these results in the value 37500 obtained by thin-layer gel chromatography in 6 M guanidine-HCl. Sedimentation velocity experiments revealed a sedimentation coefficient s20,w equals 3.4 S. The Stokes radius was 2.77 nm, the partial specific volume v 0.747 ml x g(-1). The diffusion coefficient was found to be 76.9 mum2 x s(-1) by analytical gel filtration. From these data a molecular weight of 44000 was calculated. Other physical constants of bovine-liver phosphoglycerate kinase are: frictional ratio f/f0 equals 1.18, axial ratio equals 3.3, maximal degree of hydration equals 0.1 g per g of protein. Bovine-layer phosphoglycerate kinase could not be dissociated into smaller subunits by treatments which have caused dissociation of various other proteins (8 M urea, 6 M guanidine-HCl, dodecyl sulfate, carboxymethylation, maleylation). All experiments strongly support the lack of subunit structure of the enzyme. Some characteristics of bovine-liver phosphoglycerate kinase are compared with the corresponding proteins from rabbit muscle, yeast and human erythrocytes.  相似文献   

16.
Previous work has shown that the internal pH of dormant spores of Bacillus species is more than 1 pH U below that of growing cells but rises to that of growing cells in the first minutes of spore germination. In the present work the internal pH of the whole Bacillus megaterium sporangium was measured by the distribution of the weak base methylamine and was found to decrease by approximately 0.4 during sporulation. By using fluorescence ratio image analysis with a fluorescein derivative, 2',7'-bis(2-carboxyethyl)-5 (and -6)-carboxyfluorescein (BCECF), whose fluorescence is pH sensitive, the internal pH of the mother cell was found to remain constant during sporulation at a value of 8.1, similar to that in the vegetative cell. Whereas the internal pH of the forespore was initially approximately 8.1, this value fell to approximately 7.0 approximately 90 min before synthesis of dipicolinic acid and well before accumulation of the depot of 3-phosphoglyceric acid. The pH in the forespore compartment was brought to that of the mother cell by suspending sporulating cells in a pH 8 potassium phosphate buffer plus the ionophore nigericin to clamp the internal pH of the cells to that of the external medium. We suggest that at a minimum, acidification of the forespore may regulate the activity of phosphoglycerate mutase, which is the enzyme known to be regulated to allow 3-phosphoglyceric acid accumulation during sporulation.  相似文献   

17.
1,2-Dihydroxynaphthalene oxygenase was purified from Pseudomonas putida NCIB 9816 grown on naphthalene as the sole source of carbon and energy. The enzyme had a subunit molecular weight of 19,000 and in a medium containing phosphate buffer, 1 mM mercaptoethanol, and 10% (vol/vol) ethanol had a native molecular weight greater than 275,000. The enzyme required Fe2+ for activity. It was inactivated slowly on standing, and inactivation was accelerated by dilution with aerated buffers and by H2O2. Bathophenanthroline sulfonate, o-phenanthroline, 8-hydroxyquinoline, and 2,2'-dipyridyl also inhibited the enzyme. The inactive enzyme was reactivated by anaerobic incubation with ferrous sulfate and ferrous ammonium sulfate. Thiol reagents and acetone, ethanol, or glycerol decreased the rate of loss of activity. The enzyme was most active with 1,2-dihydroxynaphthalene, for which the Km was 2.8 X 10(-4) M. 3-Methyl- and 4-methylcatechols were oxidized at 3 and 1.5%, respectively, of the rate of 1,2-dihydroxynaphthalene, and the Km for 3-methylcatechol was 1.5 X 10(-4) M. Purified 1,2-dihydroxynaphthalene oxygenase catalyzed the oxidation of 1,2-dihydroxynaphthalene, leading to the appearance of 2-hydroxychromene-2-carboxylic acid, but 3-methylcatechol was oxidized by this enzyme to 2-hydroxy-6-oxoheptadienoic acid. Thus, a product structurally analogous to 2-hydroxychromene-2-carboxylic acid was not observed when 3-methylcatechol was oxidized. This may indicate that 2-hydroxychromene-2-carboxylic acid results from cyclization of a ring fission product before release from the enzyme.  相似文献   

18.
Trimethylamine dehydrogenase (TMADH) and dimethylamine dehydrogenase (DMADH) were purified from Hyphomicrobium X. The absorbance spectra of the two enzymes were similar with λmax = 443 nm for TMADH and 440 nm for DMADH. DMADH had an apparent molecular weight of 138,000 daltons and was composed of two subunits of similar molecular weights. DMADH contained 3.91 atoms S and 4.55 atoms Fe per mole of the enzyme. Both DMADH and TMADH contained a covalently bound yellow coenzyme. The coenzyme-peptides obtained from DMADH and TMADH of Hyphomicrobium X by tryptic-chymotryptic digestion were partially purified and found to differ electrophoretically and chromatographically from the coenzyme-peptide obtained similarly from TMADH of bacterium W3A1. After digestion with aminopeptidase M the aminoacyl-coenzymes from the three enzymes had identical spectral, electrophoretic and chromatographic properties. DMADH is only the second enzyme yet found to contain 6-S-cysteinyl-FMN as coenzyme. Dissimilarities between the coenzyme-peptides of DMADH and TMADH from either Hyphomicrobium X or bacterium W3A1 are consequently located in the peptide component.  相似文献   

19.
An esterase that deacetylates cephalosporins was recovered from the supernatant of a Bacillus subtilis culture. It was partially purified by ammonium sulfate fractionation and ultrafiltration. The enzyme had a temperature optimum between 40 and 50 C and a pH optimum of 7.0. The molecular weight was estimated by gel filtration to be 190,000. The enzyme was very stable and retained greater than 80% of its activity after storage in solution at 25 C for 1 month. The esterase exhibited Michaelis-Menton kinetics with the substrates 7-aminocephalosporanic acid (7-ACA) and 7-(thiophene-2-acetamido)cephalosporanic acid (cephalothin); the K(m) values were 2.8 X 10(-3) and 8.3 X 10(-3) M, respectively. The products of 7-ACA deacetylation were weak competitive inhibitors, and a K(i) value of 5.0 X 10(-2) M was determined for acetate and of 3.6 X 10-2 M for deacetyl-7-ACA. Weak product inhibition did not prevent the deacetylation reaction from going to completion. A 5-mg/ml solution of partially purified esterase completely hydrolyzed (greater than 99.5%) a 24-mg/ml solution of 7-ACA in 3 h. Because of the kinetic properties and excellent stability, this enzyme may be useful in an immobilized form to prepare large quantities of deacetylated cephalosporin derivatives.  相似文献   

20.
1,2-Dihydroxynaphthalene dioxygenase was purified to homogeneity from a bacterium that degrades naphthalenesulfonic acids (strain BN6). The enzyme requires Fe2+ for maximal activity and consists of eight identical subunits with a molecular weight of about 33,000. Analysis of the NH2-terminal amino acid sequence revealed a high degree of homology (22 of 29 amino acids) with the NH2-terminal amino acid sequence of 2,3-dihydroxybiphenyl dioxygenase from strain Pseudomonas paucimobilis Q1. 1,2-Dihydroxynaphthalene dioxygenase from strain BN6 shows a wide substrate specificity and also cleaves 5-, 6-, and 7-hydroxy-1,2-dihydroxynaphthalene, 2,3- and 3,4-dihydroxybiphenyl, catechol, and 3-methyl- and 4-methylcatechol. Similar activities against the hydroxy-1,2-dihydroxynaphthalenes were also found in cell extracts from naphthalene-degrading bacteria.  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