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.
Aminopropionaldehyde dehydrogenase was purified to apparent homogeneity from 1,3-diaminopropane-grown cells of Arthrobacter sp. TMP-1. The native molecular mass and the subunit molecular mass of the enzyme were approximately 20,5000 and 52,000, respectively, suggesting that the enzyme is a tetramer of identical subunits. The apparent Michaelis constant (K(m)) for 1,3-diaminopropane was approximately 3 microM. The enzyme equally used both NAD(+) and NADP(+) as coenzymes. The apparent K(m) values for NAD(+) and NADP(+) were 255 microM and 108 microM, respectively. The maximum reaction rates (V(max)) for NAD(+) and NADP(+) were 102 and 83.3 micromol min(-1) mg(-1), respectively. Some tested aliphatic aldehydes and aromatic aldehydes were inert as substrates. The optimum pH was 8.0-8.5. The enzyme was sensitive to sulfhydryl group-modifying reagents.  相似文献   

2.
Protocatechuate (PCA) is the key intermediate metabolite in the lignin degradation pathway of Sphingomonas paucimobilis SYK-6 and is metabolized to pyruvate and oxaloacetate via the PCA 4,5-cleavage pathway. We characterized the 4-carboxy-2-hydroxymuconate-6-semialdehyde (CHMS) dehydrogenase gene (ligC). CHMS is the 4,5-cleavage product of PCA and is converted into 2-pyrone-4,6-dicarboxylate (PDC) by LigC. We found that ligC was located 295 bp downstream of ligB, which encodes the large subunit of the PCA 4,5-dioxygenase. The ligC gene consists of a 945-bp open reading frame encoding a polypeptide with a molecular mass of 34,590 Da. The deduced amino acid sequence of ligC showed 19 to 20% identity with 3-chlorobenzoate cis-dihydrodiol dehydrogenase of Alcaligenes sp. strain BR60 and phthalate cis-dihydrodiol dehydrogenases of Pseudomonas putida NMH102-2 and Burkholderia cepacia DBO1, which are unrelated to group I, II, and III microbial alcohol dehydrogenases (M. F. Reid and C. A. Fewson, Crit. Rev. Microbiol. 20:13-56, 1994). The ligC gene was expressed in Escherichia coli and LigC was purified to near homogeneity. Production of PDC from CHMS catalyzed by LigC was confirmed in the presence of NADP(+) by electrospray ionization-mass spectrometry and gas chromatography-mass spectrometry. LigC is a homodimer. The isoelectric point, optimum pH, and optimum temperature were estimated to be 5.3, 8.0, and 25 degrees C, respectively. The K(m) for NADP(+) was estimated to be 24.6 +/- 1.5 microM, which was approximately 10 times lower than that for NAD(+) (252 +/- 3.9 microM). The K(m)s for CHMS in the presence of NADP(+) and NAD(+) are 26.0 +/- 0.5 and 20.6 +/- 1.0 microM, respectively. Disruption of ligC in S. paucimobilis SYK-6 prevented growth with vanillate. Only PCA was accumulated during the incubation of vanillate with the whole cells of the ligC insertion mutant (DLC), indicating a lack of PCA 4,5-dioxygenase activity in DLC. However, the introduction of ligC into DLC restored its ability to grow on vanillate. PDC was suggested to be an inducer for ligAB gene expression.  相似文献   

3.
Histamine dehydrogenase (NSHADH) can be isolated from cultures of Nocardioides simplex grown with histamine as the sole nitrogen source. A previous report suggested that NSHADH might contain the quinone cofactor tryptophan tryptophyl quinone (TTQ). Here, the hdh gene encoding NSHADH is cloned from the genomic DNA of N. simplex, and the isolated enzyme is subjected to a full spectroscopic characterization. Protein sequence alignment shows NSHADH to be related to trimethylamine dehydrogenase (TMADH: EC 1.5.99.7), where the latter contains a bacterial ferredoxin-type [4Fe-4S] cluster and 6-S-cysteinyl FMN cofactor. NSHADH has no sequence similarity to any TTQ containing amine dehydrogenases. NSHADH contains 3.6+/-0.3 mol Fe and 3.7+/-0.2 mol acid labile S per subunit. A comparison of the UV/vis spectra of NSHADH and TMADH shows significant similarity. The EPR spectrum of histamine reduced NSHADH also supports the presence of the flavin and [4Fe-4S] cofactors. Importantly, we show that NSHADH has a narrow substrate specificity, oxidizing only histamine (K(m)=31+/-11 microM, k(cat)/K(m)=2.1 (+/-0.4)x10(5)M(-1)s(-1)), agmatine (K(m)=37+/-6 microM, k(cat)/K(m)=6.0 (+/-0.6)x10(4)M(-1)s(-1)), and putrescine (K(m)=1280+/-240 microM, k(cat)/K(m)=1500+/-200 M(-1)s(-1)). A kinetic characterization of the oxidative deamination of histamine by NSHADH is presented that includes the pH dependence of k(cat)/K(m) (histamine) and the measurement of a substrate deuterium isotope effect, (D)(k(cat)/K(m) (histamine))=7.0+/-1.8 at pH 8.5. k(cat) is also pH dependent and has a reduced substrate deuterium isotope of (D)(k(cat))=1.3+/-0.2.  相似文献   

4.
Flavocytochrome c-sulfide dehydrogenases (FCSDs) are complexes of a flavoprotein with a c-type cytochrome performing hydrogen sulfide-dependent cytochrome c reduction in vitro. The amino acid sequence analysis revealed that the phylogenetic relationship of different flavoproteins reflected the relationship of sulfur-oxidizing bacteria. The flavoprotein SoxF of Paracoccus pantotrophus is 29-67% identical to the flavoprotein subunit of FCSD of phototrophic sulfur-oxidizing bacteria. Purification of SoxF yielded a homogeneous emerald-green monomeric protein of 42 797 Da. SoxF catalyzed sulfide-dependent horse heart cytochrome c reduction at the optimum pH of 6.0 with a k(cat) of 3.9 s(-1), a K(m) of 2.3 microM for sulfide, and a K(m) of 116 microM for cytochrome c, as determined by nonlinear regression analysis. The yield of 1.9 mol of cytochrome c reduced per mole of sulfide suggests sulfur or polysulfide as the product. Sulfide dehydrogenase activity of SoxF was inhibited by sulfur (K(i) = 1.3 microM) and inactivated by sulfite. Cyanide (1 mM) inhibited SoxF activity at pH 6.0 by 25% and at pH 8.0 by 92%. Redox titrations in the infrared spectral range from 1800 to 1200 cm(-1) and in the visible spectral range from 400 to 700 nm both yielded a midpoint potential for SoxF of -555 +/- 10 mV versus Ag/AgCl at pH 7.5 and -440 +/- 20 mV versus Ag/AgCl at pH 6.0 (-232 mV versus SHE') and a transfer of 1.9 electrons. Electrochemically induced FTIR difference spectra of SoxF as compared to those of free flavin in solution suggested a strong cofactor interaction with the apoprotein. Furthermore, an activation/variation of SoxF during the redox cycles is observed. This is the first report of a monomeric flavoprotein with sulfide dehydrogenase activity.  相似文献   

5.
Random sequencing of a peppermint essential oil gland secretory cell cDNA library revealed a large number of clones that specified redox-type enzymes. Full-length acquisitions of each type were screened by functional expression in Escherichia coli using a newly developed in situ assay. cDNA clones encoding the monoterpene double-bond reductases (-)-isopiperitenone reductase and (+)-pulegone reductase were isolated, representing two central steps in the biosynthesis of (-)-menthol, the principal component of peppermint essential oil, and the first reductase genes of terpenoid metabolism to be described. The (-)-isopiperitenone reductase cDNA has an open reading frame of 942 nucleotides that encodes a 314 residue protein with a calculated molecular weight of 34,409. The recombinant reductase has an optimum pH of 5.5, and K(m) values of 1.0 and 2.2 microM for (-)-isopiperitenone and NADPH, respectively, with k(cat) of 1.3s(-1) for the formation of the product (+)-cis-isopulegone. The (+)-pulegone reductase cDNA has an open reading frame of 1026 nucleotides and encodes a 342 residue protein with a calculated molecular weight of 37,914. This recombinant reductase catalyzes the reduction of the 4(8)-double bond of (+)-pulegone to produce both (-)-menthone and (+)-isomenthone in a 55:45 ratio, has an optimum pH of 5.0, and K(m) values of 2.3 and 6.9 microM for (+)-pulegone and NADPH, respectively, with k(cat) of 1.8s(-1). Deduced sequence comparison revealed that these two highly substrate specific double-bond reductases show less than 12% identity. (-)-Isopiperitenone reductase is a member of the short-chain dehydrogenase/reductase superfamily and (+)-pulegone reductase is a member of the medium-chain dehydrogenase/reductase superfamily, implying very different evolutionary origins in spite of the similarity in substrates utilized and reactions catalyzed.  相似文献   

6.
Lobo S  Florova G  Reynolds KA 《Biochemistry》2001,40(39):11955-11964
Acetyl-CoA:acyl carrier protein (ACP) transacylase (ACT) activity has been demonstrated for the 3-ketoacyl-ACP synthase III (KASIII) which initiates fatty acid biosynthesis in the type II dissociable fatty acid synthases of plants and bacteria. Several lines of evidence have indicated the possibility of ACT activity being associated with proteins other than KASIII. Using a crude extract of Streptomyces collinus, we have resolved from KASIII an additional protein with ACT activity and subsequently purified it 85-fold in five chromatographic steps. The 45 kDa protein was shown by gel filtration to have a molecular mass of 185 +/- 35 kDa, consistent with a homotetrameric structure for the native enzyme. The corresponding gene (fadA) was cloned and sequenced and shown to encode a protein with amino acid sequence homology to type II thiolases. The fadA was expressed in Escherichia coli, and the resulting recombinant FadA enzyme purified by metal chelate chromatography was shown to have both ACT and thiolase activities. Kinetic studies revealed that in an ACT assay FadA had a substrate specificity for a two-carbon acetyl-CoA substrate (K(m) 8.7 +/- 1.4 microM) but was able to use ACPs from both type II fatty acid and polyketide synthases (Streptomyces glaucescens FabC ACP, K(m) 10.7 +/- 1.4 microM; E. coli FabC ACP, K(m) 8.8 +/- 2 microM; FrenN ACP, K(m) 44 +/- 12 microM). In the thiolase assay kinetic analyses revealed similar K(m) values for binding of substrates acetoacetyl-CoA (K(m) 9.8 +/- 0.8 microM) and CoA (K(m) 10.9 +/- 1.8 microM). A Cys92Ser mutant of FadA possessed virtually unchanged K(m) values for acetoacetyl-CoA and CoA but had a greater than 99% decrease in k(cat) for the thiolase activity. No detectable ACT activity was observed for the Cys92Ser mutant, demonstrating that both activities are associated with FadA and likely involve formation of the same covalent acetyl-S-Cys enzyme intermediate. An ACT activity with ACP has not previously been observed for thiolases and in the case of the S. collinus FadA is significantly lower (k(cat) 3 min(-1)) than the thiolase activity of FadA (k(cat) 2170 min(-1)). The ACT activity of FadA is comparable to the KAS activity and significantly higher than the ACT activity, reported for a streptomycete KASIII.  相似文献   

7.
In this study two open reading frames, namely HP0044 and HP0045 from H. pylori, were cloned and overexpressed in E. coli. The two recombinant proteins were demonstrated to have GDP-d-mannose 4,6-dehydratase (GMD) and GDP-l-fucose synthetase (GFS) activities, respectively. The recombinant GMD was a tetramer and had an optimum pH of 6.5. Exogenous NADP(+) was essential for its activity. The K(m) and K(cat) for GDP-d-mannose were 117.3 microM and 0.27 s(-1), respectively. The recombinant GFS was a homodimer with an optimum pH of 8.0. The K(m) and K(cat) for GDP-4-keto-6-deoxy-d-mannose were 64.08 microM and 0.75 s(-1), respectively. It can use both NADPH and NADH, but less efficient with the latter. Amino acid sequence alignment and phylogenetic analysis showed that H. pylori GFS was highly homologous to the GFS of E. coli O111 and both of them were located on a separate phylogenetic branch from other GFS. The unique clustering and origin of the two genes were also discussed.  相似文献   

8.
Zhang H  Zhou Y  Bao H  Liu HW 《Biochemistry》2006,45(26):8163-8173
Vi antigen, the virulence factor of Salmonella typhi, has been used clinically as a molecular vaccine. TviB and TviC are two enzymes involved in the formation of Vi antigen, a linear polymer consisting of alpha-1,4-linked N-acetylgalactosaminuronate. Protein sequence analysis suggests that TviB is a dehydrogenase and TviC is an epimerase. Both enzymes are expected to be NAD(+) dependent. In order to verify their functions, TviB and TviC were cloned, expressed in Escherichia coli, and characterized. The C-terminal His(6)-tagged TviB protein, purified from soluble cell fractions in the presence of 10 mM DTT, shows UDP-N-acetylglucosamine 6-dehydrogenase activity and is capable of catalyzing the conversion of UDP-N-acetylglucosamine (UDP-GlcNAc) to UDP-N-acetylglucosaminuronic acid (UDP-GlcNAcA) with a k(cat) value of 15.5 +/- 1.0 min(-)(1). The K(m) values of TviB for UDP-GlcNAc and NAD(+) are 77 +/- 9 microM and 276 +/- 52 microM, respectively. TviC, purified as C-terminal hexahistidine-tagged protein, shows UDP-GlcNAcA 4-epimerase and UDP-N-acetylgalactosamine (UDP-GalNAc) 4-epimerase activities. The K(m) values of TviC for UDP-GlcNAcA and UDP-N-acetylgalactosaminuronic acid (UDP-GalNAcA) are 20 +/- 1 microM and 42 +/- 2 microM, respectively. The k(cat) value for the conversion of UDP-GlcNAcA to UDP-GalNAcA is 56.8 +/- 0.5 min(-)(1), while that for the reverse reaction is 39.1 +/- 0.6 min(-)(1). These results show that the biosynthesis of Vi antigen is initiated by the TviB-catalyzed oxidation of UDP-GlcNAc to UDP-GalNAc, followed by the TviC-catalyzed epimerization at C-4 to form UDP-GalNAcA, which serves as the building block for the formation of Vi polymer. These results set the stage for future in vitro biosynthesis of Vi antigen. These enzymes may also be drug targets to inhibit Vi antigen production.  相似文献   

9.
Glutaconyl-coenzyme A (CoA) is the presumed enzyme-bound intermediate in the oxidative decarboxylation of glutaryl-CoA that is catalyzed by glutaryl-CoA dehydrogenase. We demonstrated glutaconyl-CoA bound to glutaryl-CoA dehydrogenase after anaerobic reduction of the dehydrogenase with glutaryl-CoA. Glutaryl-CoA dehydrogenase also has intrinsic enoyl-CoA hydratase activity, a property of other members of the acyl-CoA dehydrogenase family. The enzyme rapidly hydrates glutaconyl-CoA at pH 7.6 with a k(cat) of 2.7 s(-1). The k(cat) in the overall oxidation-decarboxylation reaction at pH 7.6 is about 9 s(-1). The binding of glutaconyl-CoA was quantitatively assessed from the K(m) in the hydratase reaction, 3 microM, and the K(i), 1.0 microM, as a competitive inhibitor of the dehydrogenase. These values compare with K(m) and K(i) of 4.0 and 12.9 microM, respectively, for crotonyl-CoA. Glu370 is the general base catalyst in the dehydrogenase that abstracts an alpha-proton of the substrate to initiate the catalytic pathway. The mutant dehydrogenase, Glu370Gln, is inactive in the dehydrogenation and the hydratase reactions. However, this mutant dehydrogenase decarboxylates glutaconyl-CoA to crotonyl-CoA without oxidation-reduction reactions of the dehydrogenase flavin. Addition of glutaconyl-CoA to this mutant dehydrogenase results in a rapid, transient increase in long-wavelength absorbance (lambda(max) approximately 725 nm), and crotonyl-CoA is found as the sole product. We propose that this 725 nm-absorbing species is the delocalized crotonyl-CoA anion that follows decarboxylation and that the decay is the result of slow protonation of the anion in the absence of the general acid catalyst, Glu370(H(+)). In the absence of detectable oxidation-reduction, the data indicate that oxidation-reduction of the dehydrogenase flavin is not essential for decarboxylation of glutaconyl-CoA.  相似文献   

10.
A new xylanase gene, xynAS9, was cloned from Streptomyces sp. S9, which was isolated from Turpan Basin, China. The full-length gene consists of 1,395 bp and encodes 465 amino acids including 38 residues of a putative signal peptide. The overall amino acid sequence shares the highest identity (50.8%) with a putative endo-1,4-beta-xylanase from Streptomyces avermitilis of the glycoside hydrolase family 10. The gene fragment encoding the mature xylanase was expressed in Escherichia coli BL21 (DE3). The recombinant protein was purified to electrophoretic homogeneity and subsequently characterized. The optimal pH and temperature for the recombinant enzyme were 6.5 and 60 degrees C, respectively. The enzyme showed broad temperature adaptability, retaining more than 65% of the maximum activity when assayed at 50-80 degrees C. The enzyme also had good thermal and pH stability. The K (m) values for oat spelt xylan and birchwood xylan substrates were 2.85 and 2.43 mg ml(-1), with the V (max) values of 772.20 and 490.87 mumol min(-1) mg(-1), respectively. The hydrolysis products of xylan were mainly xylose and xylobiose. These favorable properties should make XynAS9 a good candidate in various industrial applications.  相似文献   

11.
A nonsporulating fungus isolated from dioxine-containing tropical soils forms cellobiose dehydrogenase, when grown in media supplemented by a source of cellulose. The enzyme purified to homogeneity by SDS-PAGE (yield, 43%) had an M(r) of 95 kDa; its pH optimum was in the range 5.5-7.0; more than 50% activity was retained at pH 4.0-8.0 (citrate-phosphate buffer). The absorption spectrum of the enzyme in the visible range had the characteristic appearance of flavocytochrome proteins. Cellobiose dehydrogenase oxidized cellobiose and lactose (the respective K(M) values at pH 6.0 equaled 4.5 +/- 1.5 and 56 microM) in the presence of dichlorophenolindophenol (K(M) app = 15 +/- 3 microM at pH 6.0) taken as an electron acceptor. Other sugars were barely if at all oxidized by the enzyme. Neither ethyl-beta-D-cellobioside, heptobiose, nor chitotriose inhibited the enzymatic oxidation of lactose, even under the conditions of 100-fold molar excess. The enzyme was weakly inhibited by sodium azide dichlorophenolindophenol reduction and exhibited affinity to amorphous cellulose. At 55 degrees C and pH 6.0 (optimum stability), time to half-maximum inactivation equaled 99 min. The enzyme reduced by cellobiose was more stable than the nonreduced form. Conversely, the presence of an oxidizer (dichlorophenolindophenol) decreased the stability eight times at pH 6.0. In addition, the enzyme acted as a potent reducer of the single-electron acceptor cytochrome c3+ (K(M) app = 15 microM at pH 6.0).  相似文献   

12.
The catalytic mechanism underlying the aminopeptidase from Streptomyces griseus (SGAP) was investigated. pH-dependent activity profiles revealed the enthalpy of ionization for the hydrolysis of leucine-para-nitroanilide by SGAP. The value obtained (30 +/- 5 kJ.mol(-1)) is typical of a zinc-bound water molecule, suggesting that the zinc-bound water/hydroxide molecule acts as the reaction nucleophile. Fluoride was found to act as a pure noncompetitive inhibitor of SGAP at pH values of 5.9-8 with a K(i) of 11.4 mM at pH 8.0, indicating that the fluoride ion interacts equally with the free enzyme as with the enzyme-substrate complex. pH-dependent pK(i) experiments resulted in a pK(a) value of 7.0, suggesting a single deprotonation step of the catalytic water molecule to an hydroxide ion. The number of proton transfers during the catalytic pathway was determined by monitoring the solvent isotope effect on SGAP and its general acid-base mutant SGAP(E131D) at different pHs. The results indicate that a single proton transfer is involved in catalysis at pH 8.0, whereas two proton transfers are implicated at pH 6.5. The role of Glu131 in binding and catalysis was assessed by determining the catalytic constants (K(m), k(cat)) over a temperature range of 293-329 degrees K for both SGAP and the E131D mutant. For the binding step, the measured and calculated thermodynamic parameters for the reaction (free energy, enthalpy and entropy) for both SGAP and the E131D mutant were similar. By contrast, the E131D point mutation resulted in a four orders of magnitude decrease in k(cat), corresponding to an increase of 9 kJ.mol(-1) in the activation energy for the E131D mutant, emphasizing the crucial role of Glu131 in catalysis.  相似文献   

13.
A xylanase-coding gene (xynAHJ3, 1,104 bp) was cloned from Lechevalieria sp. HJ3 harbored in a saline soil sampled from Heijing town, aka the "town of salt", on the famous "Silk Route of the South". The gene encodes a 367-residue polypeptide (XynAHJ3) with the highest identity of 74.0 % with the endoxylanase from Streptomyces thermocarboxydus HY-15. The coding sequence of the mature protein (without the predicted signal peptide from M1 to S22) of xynAHJ3 was expressed in Escherichia coli BL21 (DE3). The activity of the purified recombinant XynAHJ3 (rXynAHJ3) was apparently optimal at 70 °C and pH 6.0, retained greater than 55 % xylanase activity at a concentration of 0.2-2.0 M Na(+) and 26 % at 4.0 M Na(+) (pH 7.5 20 °C), and showed 110.2 and 44.2 % xylanase activities in the presence of 100 mM SDS (pH 6.0 37 °C) and 10 % ethanol (pH 5.0 37 °C), respectively. rXynAHJ3 activity was stable at 50 °C and pH 4.0-11.0 for more than 60 min, in trypsin or proteinase K at 20 °C for 24 h (pH 7.5), in 10 % ethanol (v/v) (pH 5.0) at 30 or 37 °C for 72 h, in 80 % ethanol (v/v) for 1 h, and in 0.6 or 3 M NaCl (20 °C, pH 7.5) for 72 h. Compared with the majority of xylanases with tolerance to ethanol, salt, SDS, or protease (K (m) values of 1.42-15.1 mg ml(-1)), rXynAHJ3 showed a low K (m) value (0.8 mg ml(-1)) and showed only limited amino acid sequence identity with those other xylanases (less than 47 %).  相似文献   

14.
A soluble (100,000 x g supernatant) methyltransferase catalyzing the transfer of the methyl group of S-adenosyl-L-methionine to catechols was present in cell extracts of Streptomyces griseus. A simple, general, and rapid catechol-based assay method was devised for enzyme purification and characterization. The enzyme was purified 141-fold by precipitation with ammonium sulfate and successive chromatography over columns of DEAE-cellulose, DEAE-Sepharose, and Sephacryl S-200. The purified cytoplasmic enzyme required 10 mM magnesium for maximal activity and was catalytically optimal at pH 7. 5 and 35 degrees C. The methyltransferase had an apparent molecular mass of 36 kDa for both the native and denatured protein, with a pI of 4.4. Novel N-terminal and internal amino acid sequences were determined as DFVLDNEGNPLENNGGYXYI and RPDFXLEPPYTGPXKARIIRYFY, respectively. For this enzyme, the K(m) for 6,7-dihydroxycoumarin was 500 +/- 21.5 microM, and that for S-adenosyl-L-methionine was 600 +/- 32.5 microM. Catechol, caffeic acid, and 4-nitrocatechol were methyltransferase substrates. Homocysteine was a competitive inhibitor of S-adenosyl-L-methionine, with a K(i) of 224 +/- 20.6 microM. Sinefungin and S-adenosylhomocysteine inhibited methylation, and the enzyme was inactivated by Hg(2+), p-chloromercuribenzoic acid, and N-ethylmaleimide.  相似文献   

15.
Mammalian dihydroorotate dehydrogenase, the fourth enzyme of pyrimidine de novo synthesis is an integral protein of the inner mitochondrial membrane that faces the intermembrane space and is functionally connected to the respiratory chain via ubiquinone. Here, we describe the first cloning and analyzing of the complete cDNA of mouse dihydroorotate dehydrogenase. Based on our recent functional expression of the full-length rat and human dihydroorotate dehydrogenase, here we expressed N-terminal-truncated C-terminal-histidine-tagged constructs of the mouse, rat and human enzymes in Escherichia coli. These proteins were devoid of the N-terminal bipartite sequence consisting of the mitochondrial targeting sequence and adjacent hydrophobic domain necessary for import and proper location and fixation of the enzyme in the inner mitochondrial membrane. By employing metal-chelate affinity chromatography under native conditions, the enzymes were purified without detergents to a specific activity of more than 100 micromol x min(-1) x mg(-1) at pH optimum of 8.0--8.1. Flavin analyses by UV-visible spectrometry of the native enzymes gave fairly stoichiometric ratios of 0.6--1.2 mol flavin per mol protein. The kinetic constants of the truncated rat enzyme (K(m) = 11 microM dihydroorotate; K(m) = 7 microM ubiquinone) and human enzyme (K(m) = 10 microM dihydroorotate; K(m) = 14 microM ubiquinone) were very close to those recently reported for the full-size enzymes. The constants for the mouse enzyme, K(m) = 26 microM dihydroorotate and K(m) = 62 microM ubiquinone, were slightly elevated in comparison to those of the other species. The three truncated enzymes were tested for their efficacy with five inhibitors of topical clinical relevance against autoimmune disorders and tumors. Whereas the presence of the N-terminus of dihydroorotate dehydrogenase was essentially irrelevant for the efficacy of the malononitrilamides A77-1726, MNA715 and MNA279 with the rat and human enzyme, the N-termini were found to be important for the efficacy of the dianisidine derivative redoxal. Moreover, the complete N-terminal part of the human enzyme seemed to be of crucial importance for the 'slow-binding' features of the cinchoninic acid derivative brequinar, which was suggested to be one of the reasons for the narrow therapeutic window reported from clinical trials on its anti-proliferative and immunosuppressive action.  相似文献   

16.
Ge L  Seah SY 《Journal of bacteriology》2006,188(20):7205-7210
Pseudomonas aeruginosa is an opportunistic pathogen that produces the siderophore pyoverdine, which enables it to acquire the essential nutrient iron from its host. Formation of the iron-chelating hydroxamate functional group in pyoverdine requires the enzyme PvdA, a flavin-dependent monooxygenase that catalyzes the N(5) hydroxylation of l-ornithine. pvdA from P. aeruginosa was successfully overexpressed in Escherichia coli, and the enzyme was purified for the first time. The enzyme possessed its maximum activity at pH 8.0. In the absence of l-ornithine, PvdA has an NADPH oxidase activity of 0.24 +/- 0.02 micromol min(-1) mg(-1). The substrate l-ornithine stimulated this activity by a factor of 5, and the reaction was tightly coupled to the formation of hydroxylamine. The enzyme is specific for NADPH and flavin adenine dinucleotide (FAD(+)) as cofactors, as it cannot utilize NADH and flavin mononucleotide. By fluorescence titration, the dissociation constants for NADPH and FAD(+) were determined to be 105.6 +/- 6.0 microM and 9.9 +/- 0.3 microM, respectively. Steady-state kinetic analysis showed that the l-ornithine-dependent NADPH oxidation obeyed Michaelis-Menten kinetics with apparent K(m) and V(max) values of 0.58 mM and 1.34 micromol min(-1) mg(-1). l-Lysine was a nonsubstrate effector that stimulated NADPH oxidation, but uncoupling occurred and hydrogen peroxide instead of hydroxylated l-lysine was produced. l-2,4-Diaminobutyrate, l-homoserine, and 5-aminopentanoic acid were not substrates or effectors, but they were competitive inhibitors of the l-ornithine-dependent NADPH oxidation reaction, with K(ic)s of 3 to 8 mM. The results indicate that the chemical nature of effectors is important for simulation of the NADPH oxidation rate in PvdA.  相似文献   

17.
Retinol dehydrogenase 12 (RDH12) is a novel member of the short-chain dehydrogenase/reductase superfamily of proteins that was recently linked to Leber's congenital amaurosis 3 (LCA). We report the first biochemical characterization of purified human RDH12 and analysis of its expression in human tissues. RDH12 exhibits approximately 2000-fold lower K(m) values for NADP(+) and NADPH than for NAD(+) and NADH and recognizes both retinoids and lipid peroxidation products (C(9) aldehydes) as substrates. The k(cat) values of RDH12 for retinaldehydes and C(9) aldehydes are similar, but the K(m) values are, in general, lower for retinoids. The enzyme exhibits the highest catalytic efficiency for all-trans-retinal (k(cat)/K(m) approximately 900 min(-)(1) microM(-)(1)), followed by 11-cis-retinal (450 min(-)(1) mM(-)(1)) and 9-cis-retinal (100 min(-)(1) mM(-)(1)). Analysis of RDH12 activity toward retinoids in the presence of cellular retinol-binding protein (CRBP) type I or cellular retinaldehyde-binding protein (CRALBP) suggests that RDH12 utilizes the unbound forms of all-trans- and 11-cis-retinoids. As a result, the widely expressed CRBPI, which binds all-trans-retinol with much higher affinity than all-trans-retinaldehyde, restricts the oxidation of all-trans-retinol by RDH12, but has little effect on the reduction of all-trans-retinaldehyde, and CRALBP inhibits the reduction of 11-cis-retinal stronger than the oxidation of 11-cis-retinol, in accord with its higher affinity for 11-cis-retinal. Together, the tissue distribution of RDH12 and its catalytic properties suggest that, in most tissues, RDH12 primarily contributes to the reduction of all-trans-retinaldehyde; however, at saturating concentrations of peroxidic aldehydes in the cells undergoing oxidative stress, for example, photoreceptors, RDH12 might also play a role in detoxification of lipid peroxidation products.  相似文献   

18.
The steady-state kinetic properties of partially purified chicken liver sorbitol dehydrogenase (SDH) were determined spectrophotometrically at 25 degrees C, in 50 mM 3-(N-morpholino)propanesulfonic acid (MOPS) buffer, pH 8.0. In the sorbitol-to-fructose direction, analysis was based on initial rate data obtained at [NAD(+)](o)=0.1-0.4 mM and [sorbitol](o)=1.25-10 mM. The reverse process was analyzed by recording progress curves for NADH consumption, starting with [NADH](o)=0.2 mM and [fructose](o)=66.7-267 mM. The kinetics conformed to an ordered sequential model, with the cofactors adding first. The steady-state parameters in the forward direction, K(NAD(+)), K(iNAD(+)) and K(sorbitol), were found to be 210+/-62 muM, 220+/-69 microM and 3.2+/-0.54 mM, respectively. The corresponding parameters in the reverse direction were K(NADH)=240+/-58 microM, K(iNADH)=10+/-2.8 microM and K(fructose)=1000+/-140 mM. The results indicated a close parallelism with human SDH, yet up to 40-fold differences were observed when compared to related reports on other mammalian species. The structural and adaptive bases of the variation in substrate and cofactor affinities need to be accounted for.  相似文献   

19.
5-Carboxymethyl-2-hydroxymuconic semialdehyde (CHMS) dehydrogenase from Escherichia coli C and Klebsiella pneumoniae M5a1 have been purified and some of their properties studied. The apparent Km values for NAD and CHMS were 11.7 +/- 1.5 microM and 5.2 +/- 1.9 microM, respectively, for the K. pneumoniae enzyme, and 19.5 +/- 2.7 microM and 9.2 +/- 1.4 microM, respectively, for the E. coli enzyme. Both enzymes were optimally active at pH 7.5 in sodium phosphate buffer. They had subunit molecular weights of 52,000 (+/- 1000) and the native enzymes appeared to be dimers of identical subunits. The first 20 residues of their N-terminal amino acid sequences were 90% homologous. A degenerate oligonucleotide probe constructed to a six amino acid sequence common to both enzymes gave strong hybridization with DNA from E. coli strains B and W as well as with E. coli C and K. pneumoniae but little or no hybridization to DNA from E. coli K12 or Pseudomonas putida.  相似文献   

20.
Patton SM  Cropp TA  Reynolds KA 《Biochemistry》2000,39(25):7595-7604
The side chain of the antifungal polyketide ansatrienin A produced by Streptomyces collinus contains a cyclohexanecarboxylic acid (CHC) derived moiety. This CHC in the coenzyme A activated form (CHC-CoA) is derived from shikimic acid via a pathway in which the penultimate step is the isomerization of 2-cyclohexenylcarbonyl-CoA to 1-cyclohexenylcarbonyl-CoA. We have purified a 28 kDa 2-cyclohexenylcarbonyl-CoA isomerase (ChcB) from S. collinus and cloned and sequenced the corresponding chcB gene. The predicted amino acid sequence of ChcB showed moderate sequence identity to members of the hydratase/isomerase superfamily of enzymes. The recombinant ChcB was overexpressed in Escherichia coli and purified to homogeneity using metal chelate chromatography. Kinetic analysis demonstrated that recombinant ChcB had wide substrate specificity and could catalyze a double bond isomerization using 2-cyclohexenylcarbonyl-CoA (K(m) 116 +/- 68 microM, k(cat)( )()3.7 +/- 1.0 min(-)(1)), trans-3-hexenyl-CoA (K(m) 39 +/- 10 microM, k(cat)( )()12.8 +/- 1 min(-)(1)), and vinylacetyl-CoA (K(m) 156 +/- 34 microM, k(cat)( )()29 +/- 3 min(-)(1)) as substrates. ChcB activity in cell extracts of S. collinus SP1, an insertionally disrupted chcB mutant, was shown to decrease by more than 99% (as compared to the wild-type strain) using all three of these substrates. The S. collinus SP1 strain, unlike the wild-type strain, could not produce omega-cyclohexyl fatty acids but was still able to grow efficiently on methyl oleate as a sole carbon source. These observations demonstrate that the S. collinus ChcB is required for catalyzing the isomerization of 2-cyclohexenylcarbonyl-CoA to 1-cyclohexenylcarbonyl-CoA during CHC-CoA biosynthesis but not for degradation of unsaturated fatty acids. The chcB gene does not appear to be associated with the ansatrienin biosynthetic gene cluster, which has previously been shown to contain at least one gene known to be essential for CHC-CoA biosynthesis. This finding represents a notable exception to the general rule regarding the clustering of polyketide biosynthetic pathway genes.  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