首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
The NADPH-dependent dimeric flavoenzyme 4-hydroxyacetophenone monooxygenase (HAPMO) catalyzes Baeyer-Villiger oxidations of a wide range of ketones, thereby generating esters or lactones. In the current work, we probed HAPMO-coenzyme complexes present during the enzyme catalytic cycle with the aim to gain mechanistic insight. Moreover, we investigated the structural role of the nicotinamide coenzyme. For these studies, we used (i) wild type HAPMO, (ii) the R339A variant, which is active but has a low affinity toward NADPH, and (iii) the R440A variant, which is inactive but has a high affinity toward NADPH. Electrospray ionization mass spectrometry was used as the primary tool to directly observe noncovalent protein-coenzyme complexes in real time. These analyzes showed for the first time that the nicotinamide coenzyme remains bound to HAPMO during the entire catalytic cycle of the NADPH oxidase reaction. This may also have implications for other homologous Baeyer-Villiger monooxygenases. Together with the observations that NADP(+) only weakly interacts with oxidized enzyme and that HAPMO is mainly in the reduced form during catalysis, we concluded that NADP(+) interacts tightly with the reduced form of HAPMO. We also demonstrated that the association with the coenzyme is crucial for enzyme stability. The interaction with the coenzyme analog 3-aminopyridine adenine dinucleotide phosphate (AADP(+)) strongly enhanced the thermal stability of wild type HAPMO. This coenzyme-induced stabilization may also be important for related enzymes.  相似文献   

2.
Type I Baeyer–Villiger monooxygenases (BVMOs) strongly prefer NADPH over NADH as an electron donor. In order to elucidate the molecular basis for this coenzyme specificity, we have performed a site-directed mutagenesis study on phenylacetone monooxygenase (PAMO) from Thermobifida fusca. Using sequence alignments of type I BVMOs and crystal structures of PAMO and cyclohexanone monooxygenase in complex with NADP+, we identified four residues that could interact with the 2′-phosphate moiety of NADPH in PAMO. The mutagenesis study revealed that the conserved R217 is essential for binding the adenine moiety of the nicotinamide coenzyme while it also contributes to the recognition of the 2′-phosphate moiety of NADPH. The substitution of T218 did not have a strong effect on the coenzyme specificity. The H220N and H220Q mutants exhibited a ~3-fold improvement in the catalytic efficiency with NADH while the catalytic efficiency with NADPH was hardly affected. Mutating K336 did not increase the activity of PAMO with NADH, but it had a significant and beneficial effect on the enantioselectivity of Baeyer–Villiger oxidations and sulfoxidations. In conclusion, our results indicate that the function of NADPH in catalysis cannot be easily replaced by NADH. This finding is in line with the complex catalytic mechanism and the vital role of the coenzyme in BVMOs.  相似文献   

3.
Glutathione reductase from the cyanobacterium Anabaena PCC 7120 contains a pyridine-nucleotide-binding motif differing from that of the enzyme from other sources and an insertion of 10 amino acid residues. Homology modeling was used to obtain a model of the enzyme structure. It revealed that in the Anabaena enzyme Lys(203) replaces Arg, found to interact with the 2'-phosphate of NADP(H) in the enzyme from other sources, and that it has an extra loop near the entrance of the pyridine-nucleotide-binding site. The steady-state and preequilibrium kinetic properties were characterized for the wild-type enzyme, a K203R, and a loop deletion mutant. All enzyme forms had higher catalytic efficiency with NADPH than with NADH, although the difference was less than for glutathione reductase from other sources. The specificity was most pronounced in the formation of the charge-transfer complex between the pyridine nucleotide and oxidized enzyme-bound FAD, as compared to later steps in the reaction. Unexpectedly, by replacing Lys(203) with Arg, the specificity for NADPH was diminished in the complete redox reaction. Ser(174) appears to interact with the 2'-phosphate of NADPH and introduction of arginine instead of lysine, therefore, has little effect on the interaction with this coenzyme. However, the efficiency in forming the charge-transfer complex between the pyridine nucleotide and oxidized enzyme-bound FAD was increased in the K203R mutant using NADPH but not with NADH. The lack of affinity toward 2',5'-ADP-Sepharose by the wild-type enzyme was not changed by replacing Lys(203) with Arg but deletion of the loop resulted in an enzyme that bound to the immobilized ligand. Removal of the loop increased the efficiency of the enzyme in the reductive half-reaction with both pyridine-nucleotides as well as in the overall catalytic mechanism.  相似文献   

4.
A major problem when xylose is used for ethanol production is the intercellular redox imbalance arising from different coenzyme specificities of xylose reductase (XR) and xylitol dehydrogenase. The residue Lys21 in XR from Pichia stipitis was subjected to site-directed mutagenesis to alter its coenzyme specificity. The N272D mutant exhibited improved catalytic efficiency when NADH was the coenzyme. Both K21A and K21A/N272D preferred NADH to NADPH, their catalytic efficiencies for NADPH were almost zero. The catalytic efficiency of K21A/N272D for NADH was almost 9-fold and 2-fold that of K21A and the wild-type enzyme, respectively. Complete reversal of coenzyme specificity toward NADH and improved catalytic efficiency were achieved. Electronic supplementary material  The online version of this article (doi:) contains supplementary material, which is available to authorized users. Qi-Kai Zeng, Hong-Li Du, Jing-Fang Wang have contributed equally to this work.  相似文献   

5.
Siderophore A (SidA) is a flavin-dependent monooxygenase that catalyzes the NAD(P)H- and oxygen-dependent hydroxylation of ornithine in the biosynthesis of siderophores in Aspergillus fumigatus and is essential for virulence. SidA can utilize both NADPH or NADH for activity; however, the enzyme is selective for NADPH. Structural analysis shows that R279 interacts with the 2′-phosphate of NADPH. To probe the role of electrostatic interactions in coenzyme selectivity, R279 was mutated to both an alanine and a glutamate. The mutant proteins were active but highly uncoupled, oxidizing NADPH and producing hydrogen peroxide instead of hydroxylated ornithine. For wtSidA, the catalytic efficiency was 6-fold higher with NADPH as compared to NADH. For the R279A mutant the catalytic efficiency was the same with both coenyzmes, while for the R279E mutant the catalytic efficiency was 5-fold higher with NADH. The effects are mainly due to an increase in the KD values, as no major changes on the kcat or flavin reduction values were observed. Thus, the absence of a positive charge leads to no coenzyme selectivity while introduction of a negative charge leads to preference for NADH. Flavin fluorescence studies suggest altered interaction between the flavin and NADP+ in the mutant enzymes. The effects are caused by different binding modes of the coenzyme upon removal of the positive charge at position 279, as no major conformational changes were observed in the structure for R279A. The results indicate that the positive charge at position 279 is critical for tight binding of NADPH and efficient hydroxylation.  相似文献   

6.
A novel flavoprotein that catalyses the NADPH-dependent oxidation of 4-hydroxyacetophenone to 4-hydroxyphenyl acetate, was purified to homogeneity from Pseudomonas fluorescens ACB. Characterization of the purified enzyme showed that 4-hydroxyacetophenone monooxygenase (HAPMO) is a homodimer of approximately 140 kDa with each subunit containing a noncovalently bound FAD molecule. HAPMO displays a tight coupling between NADPH oxidation and substrate oxygenation. Besides 4-hydroxyacetophenone a wide range of other acetophenones are readily converted via a Baeyer-Villiger rearrangement reaction into the corresponding phenyl acetates. The P. fluorescens HAPMO gene (hapE) was characterized. It encoded a 640 amino-acid protein with a deduced mass of 71 884 Da. Except for an N-terminal extension of approximately 135 residues, the sequence of HAPMO shares significant similarity with two known types of Baeyer-Villiger monooxygenases: cyclohexanone monooxygenase (27-33% sequence identity) and steroid monooxygenase (33% sequence identity). The HAPMO sequence contains several sequence motifs indicative for the presence of two Rossman fold domains involved in FAD and NADPH binding. The functional role of a recently identified flavoprotein sequence motif (ATG) was explored by site-directed mutagenesis. Replacement of the strictly conserved glycine (G490) resulted in a dramatic effect on catalysis. From a kinetic analysis of the G490A mutant it is concluded that the observed sequence motif serves a structural function which is of importance for NADPH binding.  相似文献   

7.
p-Hydroxybenzoate hydroxylase (PHBH) is the archetype of the family of NAD(P)H-dependent flavoprotein aromatic hydroxylases. These enzymes share a conserved FAD-binding domain but lack a recognizable fold for binding the pyridine nucleotide. We have switched the coenzyme specificity of strictly NADPH-dependent PHBH from Pseudomonas fluorescens by site-directed mutagenesis. To that end, we altered the solvent exposed helix H2 region (residues 33-40) of the FAD-binding domain. Non-conservative selective replacements of Arg33 and Tyr38 weakened the binding of NADPH without disturbing the protein architecture. Introduction of a basic residue at position 34 increased the NADPH binding strength. Double (M2) and quadruple (M4) substitutions in the N-terminal part of helix H2 did not change the coenzyme specificity. By extending the replacements towards residues 38 and 40, M5 and M6 mutants were generated which were catalytically more efficient with NADH than with NADPH. It is concluded that specificity in P. fluorescens PHBH is conferred by interactions of Arg33, Tyr38 and Arg42 with the 2'-phosphate moiety of bound NADPH, and that introduction of an acidic group at position 38 potentially enables the recognition of the 2'-hydroxy group of NADH. This is the first report on the coenzyme reversion of a flavoprotein aromatic hydroxylase.  相似文献   

8.
To elucidate the structural basis for the alteration of coenzyme specificity from NADH toward NADPH in a malate dehydrogenase mutant EX7 from Thermus flavus, we determined the crystal structures at 2.0 A resolution of EX7 complexed with NADPH and NADH, respectively. In the EX7-NADPH complex, Ser42 and Ser45 form hydrogen bonds with the 2'-phosphate group of the adenine ribose of NADPH, although the adenine moiety is not seen in the electron density map. In contrast, although Ser42 and Ser45 occupy a similar position in the EX7-NADH complex structure, both the adenine and adenine ribose moieties of NADH are missing in the map. These results and kinetic analysis of site-directed mutant enzymes indicate (1) that the preference of EX7 for NADPH over NADH is ascribed to the recognition of the 2'-phosphate group by two Ser and Arg44, and (2) that the adenine moiety of NADPH is not recognized in this mutant.  相似文献   

9.
L A LeBrun  B V Plapp 《Biochemistry》1999,38(38):12387-12393
The rate of association of NAD(+) with wild-type horse liver alcohol dehydrogenase (ADH) is maximal at pH values between pK values of about 7 and 9, and the rate of NADH association is maximal at a pH below a pK of 9. The catalytic zinc-bound water, His-51 (which interacts with the 2'- and 3'-hydroxyl groups of the nicotinamide ribose of the coenzyme in the proton relay system), and Lys-228 (which interacts with the adenosine 3'-hydroxyl group and the pyrophosphate of the coenzyme) may be responsible for the observed pK values. In this study, the Lys228Arg, His51Gln, and Lys228Arg/His51Gln (to isolate the effect of the catalytic zinc-bound water) mutations were used to test the roles of the residues in coenzyme binding. The steady state kinetic constants at pH 8 for the His51Gln enzyme are similar to those for wild-type ADH. The Lys228Arg and Lys228Arg/His51Gln substitutions decrease the affinity for the coenzymes up to 16-fold, probably due to altered interactions with the arginine at position 228. As determined by transient kinetics, the rate constant for association of NAD(+) with the mutated enzymes no longer decreases at high pH. The pH profile for the Lys228Arg enzyme retains the pK value near 7. The His51Gln and Lys228Arg/His51Gln substitutions significantly decrease the rate constants for NAD(+) association, and the pH dependencies show that these enzymes bind NAD(+) most rapidly at a pH above pK values of 8. 0 and 9.0, respectively. It appears that the pK of 7 in the wild-type enzyme is shifted up by the H51Q substitutions, and the resulting pH dependence is due to the deprotonation of the catalytic zinc-bound water. Kinetic simulations suggest that isomerization of the enzyme-NAD(+) complex is substantially altered by the mutations. In contrast, the pH dependencies for NADH association with His51Gln, Lys228Arg, and Lys228Arg/His51Gln enzymes were the same as for wild-type ADH, suggesting that the binding of NAD(+) and the binding of NADH are controlled differently.  相似文献   

10.
The flavin-containing monooxygenase from Saccharomyces cerevisiae (yFMO) uses NADPH and O(2) to oxidize thiol containing substrates such as GSH and thereby generates the oxidizing potential for the ER. The enzyme uses NADPH 12 times more efficiently than NADH. Amino acid sequence analysis suggests that Lys 219 and/or Lys 227 may act as counterions to the 2' phosphate of NADPH and to help determine the preference for pyridine nucleotides. Site directed mutations show that Lys 219 makes the greater contribution to cosubstrate recognition. Conversion of Lys 219 to Ala reduces NADPH dependent activity 90-fold, but has no effect on NADH-dependent activity. Conversion of Lys 227 to Ala reduces NADPH-dependent activity fivefold and NADH-dependent activity threefold. Dissociation constants for NADP(+) to oxidized yFMO were measured spectroscopically. K(d) is 12 microM for the wild-type enzyme and 243 microM for the K219A mutant, consistent with the role of Lys 219 in pyridine nucleotide binding.  相似文献   

11.
木糖还原酶是重组酿酒酵母工程菌利用木糖生成乙醇代谢途径中的关键酶, 该关键酶在利用木糖时依赖NADPH而不是NADH是导致酿酒酵母代谢木糖生成乙醇的最终产率低的主要原因之一。为了改变树干毕赤氏酵母木糖还原酶的辅酶依赖性, 对它的第21位赖基酸Lys进行了突变。利用质粒载体pET28b分别将突变后的基因K21A-XYL1、K21R-XYL1及野生基因WT-XYL1在大肠杆菌E. coli BL21(DE3)中进行表达, 表达后的蛋白经His-Tag纯化柱纯化后测定酶学性质。结果表明: K21R突变子的辅酶依赖性没有改变, 但K21A突变子的辅酶依赖性由NADPH完全逆转为NADH。  相似文献   

12.
Carrigan JB  Engel PC 《The FEBS journal》2007,274(19):5167-5174
Glutamate dehydrogenase (EC 1.4.1.2-4) from Peptostreptococcus asaccharolyticus has a strong preference for NADH over NADPH as a coenzyme, over 1000-fold in terms of kcat/Km values. Sequence alignments across the wider family of NAD(P)-dependent dehydrogenases might suggest that this preference is mainly due to a negatively charged glutamate at position 243 (E243) in the adenine ribose-binding pocket. We have examined the possibility of altering coenzyme specificity of the Peptostreptococcus enzyme, and, more specifically, the role of residue 243 and neighbouring residues in coenzyme binding, by introducing a range of point mutations. Glutamate dehydrogenases are unusual among dehydrogenases in that NADPH-specific forms usually have aspartate at this position. However, replacement of E243 with aspartate led to only a nine-fold relaxation of the strong discrimination against NADPH. By contrast, replacement with a more positively charged lysine or arginine, as found in NADPH-dependent members of other dehydrogenase families, allows a more than 1000-fold shift toward NADPH, resulting in enzymes equally efficient with NADH or NADPH. Smaller shifts in the same direction were also observed in enzymes where a neighboring tryptophan, W244, was replaced by a smaller alanine (approximately six-fold) or Asp245 was changed to lysine (32-fold). Coenzyme binding studies confirm that the mutations result in the expected major changes in relative affinities for NADH and NADPH, and pH studies indicate that improved affinity for the extra phosphate of NADPH is the predominant reason for the increased catalytic efficiency with this coenzyme. The marked difference between the results of replacing E243 with aspartate and with positive residues implies that the mode of NADPH binding in naturally occurring NADPH-dependent glutamate dehydrogenases differs from that adopted in E243K or E243D and in other dehydrogenases.  相似文献   

13.
Jao SC  Huang LF  Hwang SM  Li WS 《Biochemistry》2006,45(6):1547-1553
Analysis of the pH-rate profile for catalysis of bradykinin cleavage by aminopeptidase P (AMPP), a manganese-containing hydrolase from Escherichia coli, was carried out to show that optimal catalytic function is obtained at neutral pH. On the basis of information derived from the crystal structure, peptidase sequence alignments, and the hydrolysis of organophosphate triesters, active site residues Arg153, Arg370, Trp88, Tyr387, and Arg404 were identified as potential catalytic residues. Site-directed mutagenesis was used to substitute these residues with Leu, Ala, Trp, Lys, or Phe. The kcat values for the Arg153, Arg370, and Trp88 mutants were nearly the same as that for the wild-type enzyme. The kcat values of the R404K, R404A, and Y387A mutants were lower by factors of 285, 400, and 16, respectively. Inductively coupled plasma mass spectrometry and circular dichroism spectroscopy showed that Arg404 is not required for metal chelation or stabilization of protein secondary structure. The hydrogen bond network observed between the side chains of conserved residues Asp260, Arg404, and Tyr387 indicated that Arg404 participates in proton relay. This was further evidenced by the return of activity in the R404A mutant by the addition of guanidine. Also, reduced catalytic efficiency in the R404K mutant, which conserves the positive charge at the bridge site, shows that only the arginine group of Arg404 (not the ammonium group of Lys404) can participate in the hydrogen bond network. The hydrogen bond interaction between the Arg404 and the Tyr387 ring hydroxyl group is suggested by the reduced catalytic efficiency of the Y387F mutant.  相似文献   

14.
It is postulated that basic residues in the regulatory region of myosin light chain kinase are important for conferring autoinhibition by binding to the catalytic core. To investigate this proposal, 10 basic amino acids within the regulatory region of rabbit smooth muscle myosin light chain kinase (Lys961-Lys979) were replaced either singularly or in combination with acidic or nonpolar residues by site-directed mutagenesis. All active mutant kinases were dependent on Ca2+/calmodulin for catalytic activity. None of the mutants was active in the absence of Ca2+/calmodulin, suggesting that the autoinhibitory region has not been defined completely. Charge reversal mutants at Arg974, Arg975, and Lys976 resulted in loss of high affinity binding of calmodulin and increased the concentration of calmodulin required for half-maximal activation (KCaM). The charge reversal mutant at Lys979 also increased KCaM but to a lesser extent. Charge reversal mutants at Lys965 and Arg967 resulted in an inactive myosin light chain kinase that could not be proteolytically activated. When these residues were mutated to Ala, the expressed kinase was dependent upon Ca2+/calmodulin for activity and exhibited a decrease in KCaM. Charge reversal mutants in Lys961 and Lys962 also had decreased KCaM values. These basic residues amino-terminal of the calmodulin binding domain may play an important role in the activation of the kinase.  相似文献   

15.
Botulinum neurotoxin (NT) is synthesized by Clostridium botulinum as about a 150-kDa single-chain polypeptide. Posttranslational modification by bacterial or exogenous proteases yielded dichain structure which formed a disulfide loop connecting a 50-kDa light chain (Lc) and 100-kDa heavy chain (Hc). We determined amino acid sequences around cleavage sites in the loop region of botulinum NTs produced by type C strain Stockholm, type D strain CB16, and type F strain Oslo by analysis of the C-terminal sequence of Lc and the N-terminal sequence of Hc. Cleavage was found at one or two sites at Arg444/Ser445 and Lys449/Thr450 for type C, and Lys442/Asn443 and Arg445/Asp446 for type D, respectively. In culture fluid of mildly proteolytic strains of type C and D, therefore, NT exists as a mixture of at least three forms of nicked dichain molecules. The NT of type F proteolytic strain Oslo showed the Arg435 as a C-terminal residue of Lc and Ala440 as an N-terminal residue of Hc, indicating that the bacterial protease cuts twice (Arg435/Lys436 and Lys439/Ala440), with excision of four amino acid residues. The location of cleavage and number of amino acid residue excisions in the loop region could be explained by the degree of exposure of amino acid residues on the surface of the molecule, which was predicted as surface probability from the amino acid sequence. In addition, the observed correlation may also be adapted to the cleavage sites of the other botulinum toxin types, A, B, E, and G.  相似文献   

16.
To identify amino acids of cytochrome P450d (P450d) which participate in the interaction with NADPH-cytochrome P450 reductase, we changed conserved ionic amino acids of P450d to others by site-directed mutagenesis. Turnover numbers (0.032-0.008 min-1) of purified mutants Lys94-Glu, Lys99-Glu, Lys105-Glu, Lys440-Glu, Lys453-Glu, Arg455-Glu, and Lys463-Glu toward 7-ethoxycoumarin were much lower than that (0.380 min-1) of the wild type at 25 degrees C. Reduction rates (less than 0.054 s-1) of the heme of all mutants (0.1 microM) in the presence of NADPH and the reductase (0.3 microM) were much lower than that (5.9 s-1) of the wild type. Furthermore, a turnover number (0.042 min-1) of a microsomal triple mutant (Arg135-Leu + Arg136-Leu + Arg137-Leu) of a conserved Arg cluster was much lower than that (0.674 min-1) of the wild type at 37 degrees C. Thus, we suggest that Lys94, Lys99, Lys105, Lys440, Lys453, Arg455, Lys463, and perhaps the Arg cluster Arg135-Arg136-Arg137 of P450d will participate in the intermolecular electron transfer process by forming ionic bridges between the two proteins and/or by orienting appropriate geometry for electron transfer on the interfacial surface between the two proteins.  相似文献   

17.
Ferredoxin-NADP(+) reductase catalyses NADP(+) reduction, being specific for NADP(+)/H. To understand coenzyme specificity determinants and coenzyme specificity reversion, mutations at the NADP(+)/H pyrophosphate binding and of the C-terminal regions have been simultaneously introduced in Anabaena FNR. The T155G/A160T/L263P/Y303S mutant was produced. The mutated enzyme presents similar k(cat) values for NADPH and NADH, around 2.5 times slower than that reported for WT FNR with NADPH. Its K(m) value for NADH decreased 20-fold with regard to WT FNR, whereas the K(m) for NADPH remains similar. The combined effect is a much higher catalytic efficiency for NAD(+)/H, with a minor decrease of that for NADP(+)/H. In the mutated enzyme, the specificity for NADPH versus NADH has been decreased from 67,500 times to only 12 times, being unable to discriminate between both coenzymes. Additionally, giving the role stated for the C-terminal Tyr in FNR, its role in the energetics of the FAD binding has been analysed.  相似文献   

18.
Nucleotide excision repair (NER) is responsible for the recognition and removal of numerous structurally unrelated DNA lesions. In prokaryotes, the proteins UvrA, UvrB and UvrC orchestrate the recognition and excision of aberrant lesions from DNA. Despite the progress we have made in understanding the NER pathway, it remains unclear how the UvrA dimer interacts with DNA to facilitate DNA damage recognition. The purpose of this study was to define amino acid residues in UvrA that provide binding energy to DNA. Based on conservation among approximately 300 UvrA sequences and 3D-modeling, two positively charged residues, Lys680 and Arg691, were predicted to be important for DNA binding. Mutagenesis and biochemical analysis of Bacillus caldontenax UvrA variant proteins containing site directed mutations at these residues demonstrate that Lys680 and Arg691 make a significant contribution toward the DNA binding affinity of UvrA. Replacing these side chains with alanine or negatively charged residues decreased UvrA binding 3-37-fold. Survival studies indicated that these mutant proteins complemented a WP2 uvrA(-) strain of bacteria 10-100% of WT UvrA levels. Further analysis by DNase I footprinting of the double UvrA mutant revealed that the UvrA DNA binding defects caused a slower rate of transfer of DNA to UvrB. Consequently, the mutants initiated the oligonucleotide incision assay nearly as well as WT UvrA thus explaining the observed mild phenotype in the survival assay. Based on our findings we propose a model of how UvrA binds to DNA.  相似文献   

19.
Formate dehydrogenase from Candida boidinii (CboFDH) catalyses the oxidation of formate anion to carbon dioxide with concomitant reduction of NAD(+) to NADH. CboFDH is highly specific to NAD(+) and virtually fails to catalyze the reaction with NADP(+). Based on structural information for CboFDH, the loop region between beta-sheet 7 and alpha-helix 10 in the dinucleotide-binding fold was predicted as a principal determinant of coenzyme specificity. Sequence alignment with other formate dehydrogenases revealed two residues (Asp195 and Tyr196) that could account for the observed coenzyme specificity. Positions 195 and 196 were subjected to two rounds of site-saturation mutagenesis and screening and enabled the identification of a double mutant Asp195Gln/Tyr196His, which showed a more than 2 x 10(7)-fold improvement in overall catalytic efficiency with NADP(+) and a more than 900-fold decrease in the efficiency with NAD(+) as cofactors. The results demonstrate that the combined polar interactions and steric factors comprise the main structural determinants responsible for coenzyme specificity. The double mutant Asp195Gln/Tyr196His was tested for practical applicability in a cofactor recycling system composed of cytochrome P450 monooxygenase from Bacillus subtilis, (CYP102A2), NADP(+), formic acid and omega-(p-nitrophenyl)dodecanoic acid (12-pNCA). Using a 1250-fold excess of 12-pNCA over NADP(+) the first order rate constant was determined to be equal to k(obs) = 0.059 +/- 0.004 min(-1).  相似文献   

20.
The 4-hydroxyacetophenone monooxygenase (HAPMO) from Pseudomonas fluorescens ACB catalyzes NADPH- and oxygen-dependent Baeyer-Villiger oxidation of 4-hydroxyacetophenone to the corresponding acetate ester. Using the purified enzyme from recombinant Escherichia coli, we found that a broad range of carbonylic compounds that are structurally more or less similar to 4-hydroxyacetophenone are also substrates for this flavin-containing monooxygenase. On the other hand, several carbonyl compounds that are substrates for other Baeyer-Villiger monooxygenases (BVMOs) are not converted by HAPMO. In addition to performing Baeyer-Villiger reactions with aromatic ketones and aldehydes, the enzyme was also able to catalyze sulfoxidation reactions by using aromatic sulfides. Furthermore, several heterocyclic and aliphatic carbonyl compounds were also readily converted by this BVMO. To probe the enantioselectivity of HAPMO, the conversion of bicyclohept-2-en-6-one and two aryl alkyl sulfides was studied. The monooxygenase preferably converted (1R,5S)-bicyclohept-2-en-6-one, with an enantiomeric ratio (E) of 20, thus enabling kinetic resolution to obtain the (1S,5R) enantiomer. Complete conversion of both enantiomers resulted in the accumulation of two regioisomeric lactones with moderate enantiomeric excess (ee) for the two lactones obtained [77% ee for (1S,5R)-2 and 34% ee for (1R,5S)-3]. Using methyl 4-tolyl sulfide and methylphenyl sulfide, we found that HAPMO is efficient and highly selective in the asymmetric formation of the corresponding (S)-sulfoxides (ee > 99%). The biocatalytic properties of HAPMO described here show the potential of this enzyme for biotechnological applications.  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