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.
A nitrilase gene from Alcaligenes sp. ECU0401 was cloned and overexpressed in Escherichia coli BL21 (DE3) in a soluble form. The encoded protein with a His6-tag was purified to nearly homogeneity as revealed by SDS-PAGE with a molecular weight of approximately 38.5 kDa, and the holoenzyme was estimated to be composed of 10 subunits of identical size by size exclusion chromatography. The V max and K m parameters were determined to be 27.9 μmol min−1 mg−1 protein and 21.8 mM, respectively, with mandelonitrile as the substrate. The purified enzyme was highly thermostable with a half life of 155 h at 30 °C and 94 h at 40 °C. Racemic mandelonitrile (50 mM) could be enantioselectively hydrolyzed to (R)-(−)-mandelic acid by the purified nitrilase with an enantiomeric excess of 97%. The extreme stability, high activity and enantioselectivity of this nitrilase provide a solid base for its practical application in the production of (R)-(−)-mandelic acid.  相似文献   

2.
Rhodococcus sp. NDB 1165, a nitrile-transforming organism was isolated from temperate forest soil of Himalayas. The nitrilase (EC 3.5.5.2) activity of this organism had higher substrate specificity toward aromatic nitriles (benzonitrile, 3-cyanopyridine and 4-cyanopyridine) and unsaturated aliphatic nitrile (acrylonitrile) in comparison to saturated aliphatic nitriles (acetonitrile, propionitrile, butyronitrile and isobutyronitrile) nitrile and arylacetonitrile (phenylacetonitrile and indole-3-acetonitrile). The nitrilase of Rhodococcus sp. NDB 1165 was inducible in nature and propionitrile proved to be an efficient inducer. However, the salts of ferrous and cobalt ions had an inhibitory effect. Under optimized reaction conditions (pH 8.0 and temperature 45°C) the nitrilase activity of this organism was 2.39 ± 0.07 U/mg dry cell mass (dcm). The half-life of this enzyme was 150 min and 40 min at 45°C and 50°C respectively. However, it was quite stable at 40°C and around 58 % activity was retained even after 6 h at this temperature. The V max and K m value of this nitrilase were 1.67 μmol/ml min and 0.1 M respectively using 3-cyanopyridine as substrate. However, the decrease in V max and K m values (0.56 μmol/ml min and 0.02 M, respectively) were ␣observed at >0.05 M 3-cyanopyridine which revealed that this enzyme experienced uncompetitive inhibition at higher substrate concentrations. Under optimized reaction conditions, 1.6 M 3-cyanopyridine was successfully converted in to nicotinic acid using 2.0 mg resting cells (dcm)/ml reaction mixture in 11 h. This is the highest production of nicotinic acid i.e. 8.95 mg/mg resting cells (dcm)/h as compared to nitrilase systems reported hitherto.  相似文献   

3.
A moderately thermophilic Gram-positive, sporulating, rod-shaped strain of Bacillus with nitrile-degrading activity was isolated from polluted industrial waters. Whole cells and cell-free extracts from the end of exponential growth phase expressed 7.6 nkat mg−1 and 2.0 nkat mg−1 benzonitrile-degrading activity, respectively, after cultivation in a fermentor with complex medium containing benzonitrile as an inducer. The benzonitrile degradation took place via the nitrilase pathway directly to benzoic acid without intermediate formation of benzamide. Samples with benzonitrilase activity of 7.6 nkat mg−1 converted 3 mg benzonitrile in 1 h at 45°C. The half-life of benzonitrilase activity for a whole cell suspension and for cells immobilized in 2% agar was 4.5 min and 6 min at 70°C without substrate and 3 min at 90°C with substrate, respectively. The nitrilase had a broad substrate spectrum. The active biocatalyst obtained by immobilization was used in a continuous process and total biodegradation of 14.1 mM benzonitrile and 37.2 mM 4-cyanopyridine in a column bioreactor at 50°C for 5 h was achieved.  相似文献   

4.
Nitrilase of Nocardia globerula NHB-2 was induced by short-chain aliphatic nitriles (valeronitrile > isobutyronitrile > butyronitrile > propionitrile) and exhibited activity towards aromatic nitriles (benzonitrile > 3-cyanopyridine > 4-cyanopyridine > m-tolunitrile > p-tolunitrile). Hyperinduction of nitrilase (6.67 U mgDCW−1, 18.7 U mL−1) was achieved in short incubation time (30 h, 30°C) through multiple feeding of isobutyronitrile in the growth medium. The nitrilase of this organism exhibits both substrate and product inhibition effects. In a fed batch reaction at 1 L scale using hyperinduced resting cells corresponding to 10 U mL−1 nitrilase activity (1.5 mgDCW mL−1), a total of 123.11 g nicotinic acid was produced at a rate of 24 g h−1 gDCW−1.  相似文献   

5.
Nitrile-hydrolyzing bacteria have the potential to perform useful biotransformations such as the production of industrially useful acids and amides. In this study, we report a nitrile-degrading bacterium with significant nitrile metabolism. Molecular characterization of 16S rDNA gene characterized this strain as Bacillus cereus. Medium optimization of B. cereus FA12 showed that biomass and nitrilase production was strongly supported by glucose (10 gL? 1) and yeast extract (10 gL? 1). Enzymatic production improved slightly in the pH range from 6.0 to 7.0. The addition of Mg+2, Fe+2, and Na+ supported biomass and nitrilase production; however, other metal ions, Co+2 and Cu+2, inhibited production. The apparent molecular mass of the puri?ed FA12 nitrilase as determined by sodium dodecyl sulfate-polyacrylamide gel electrophoresis (SDS-PAGE) was about 45 kDa. Nitrilase FA12 shows relatively high activity and stability at pH 7.0 and 40°C. Nitrilase FA12 was marginally inhibited with Ca+ 2 and Co+2, whereas inhibition in the presence of dithiothreitol or DTT was 80%. The pseudo Km (mM) values of resting cells (i.e., treating whole cells as if they were an enzyme) for acetonitrile and acetamide were determined to be 2.36 and 1.81, respectively. Under optimum situations, B. cereus FA12 resting cells produced 83 and 58 (U/mg) acetonitrile/acetamide degrading activity, respectively. Ammonia production from acetamide and acetonitrile by the B. cereus FA12 was maximum after 5 and 7 h of incubation, respectively. These results indicate that B. cereus FA12 resting cells may be used in nitrile biotransformations to produce commercially useful compounds.  相似文献   

6.
The operational stabilities of nitrilases from Aspergillus niger K10 and Fusarium solani O1 were examined with 4-cyanopyridine as the substrate in continuous-stirred membrane reactors (CSMRs). The former enzyme was fairly stable at 30 °C with a deactivation constant (k d) and enzyme half-life of 0.014 h−1 and 50 h, respectively, but the latter exhibited an even higher stability characterized by k d = 0.008 h−1 and half-life of 87 h at 40 °C. Another advantage of this enzyme was its high chemoselectivity, i.e., selective transformation of nitriles into carboxylic acids, while the amide formed a high ratio of A. niger K10 nitrilase product. High conversion rates (>90%) were maintained for about 52 h using the nitrilase from F. solani O1 immobilized in cross-linked enzyme aggregates (CLEAs). The purity of isonicotinic acid was increased from 98% to >99.9% by using two CSMRs connected in series, the first one containing the F. solani O1 nitrilase and the second the amidase from Rhodococcus erythropolis A4 (both enzymes as CLEAs), the amidase hydrolyzing the by-product isonicotinamide.  相似文献   

7.
《Process Biochemistry》2014,49(12):2141-2148
A nitrilase gene from Acidovorax facilis ZJB09122 was cloned and expressed in Escherichia coli BL21 (DE3). To improve the activity of this nitrilase, a key amino acid Phe168 was selected and mutated by site-directed mutagenesis, based on the homology modeling and previously described “hot spot” mutation. After mutation and screening, a mutant (Mut-F168V) with higher activity and stability was obtained. The nitrilase activity of Mut-F168V to hydrolyze 1-cyanocyclohexylacetonitrile was 39.52-fold compared with wild type A. facilis nitrilase (Wt-Acf-Nit). The values of Km and Vmax of Mut-F168V were markedly decreased to 1.89-fold and increased to 50.34-fold as compared to Wt-Acf-Nit, respectively. The biotransformation study showed that 1.0 M of 1-cyanocyclohexylacetonitrile could be regioselectively hydrolyzed to 1-(cyanocyclohexyl) acetic acid with 90% yield. The yield of 1-(cyanocyclohexyl) acetic acid by Mut-F168V was 66.19-fold compared to Wt-Acf-Nit after 1 h at the concentration of 1.0 M 1-cyanocyclohexylacetonitrile as substrate. The 1-(cyanocyclohexyl) acetic acid was subsequently isolated and characterized. The mutant (Mut-F168V) appears promising for potential applications for the industrial production of 1-(cyanocyclohexyl) acetic acid.  相似文献   

8.
Nitrile groups are catabolized to the corresponding acid and ammonia through one-step reaction involving a nitrilase. Here, we report the use of bioinformatic and biochemical tools to identify and characterize the nitrilase (NitPf5) from Pseudomonas fluorescens Pf-5. The nitPf5 gene was identified via sequence analysis of the whole genome of P. fluorescens Pf-5 and subsequently cloned and overexpressed in Escherichia coli. DNA sequence analysis revealed an open-reading frame of 921 bp, capable of encoding a polypeptide of 307 amino acids residues with a calculated isoelectric point of pH 5.4. The enzyme had an optimal pH and temperature of 7.0°C and 45°C, respectively, with a specific activity of 1.7 and 1.9 μmol min−1 mg protein−1 for succinonitrile and fumaronitrile, respectively. The molecular weight of the nitrilase as determined by sodium dodecyl sulfate-polyacrylamide gel electrophoresis and gel filtration chromatography was 33,000 and 138,000 Da, respectively, suggesting that the enzyme is homotetrameric. Among various nitriles, dinitriles were the preferred substrate of NitPf5 with a K m = 17.9 mM and k cat/K m = 0.5 mM−1 s−1 for succinonitrile. Homology modeling and docking studies of dinitrile and mononitrile substrate into the active site of NitPf5 shed light on the substrate specificity of NitPf5. Although nitrilases have been characterized from several other sources, P. fluorescens Pf-5 nitrilase NitPf5 is distinguished from other nitrilases by its high specific activity toward dinitriles, which make P. fluorescens NitPf5 useful for industrial applications, including enzymatic synthesis of various cyanocarboxylic acids.  相似文献   

9.
Nitrilase (E.C. 3.5.5.1) cloned from Arabidopsis thaliana converts indole-3-acetonitrile to the plant growth hormone, indole-3-acetic acid in vitro. To probe the capacity of this enzyme under physiological conditions in vivo, the cDNA PM255, encoding nitrilase II, was stably integrated into the genome of Nicotiana tabacum by direct protoplast transformation under the control of the CaMV-35S promotor. The regenerated plants appeared phenotypically normal. Nitrilase II was expressed, based on the occurrence of its mRNA and polypeptide. The enzyme was catalytically active, when extracted from leaf tissue of transgenic plants (specific activity: 25 fkat mg?1 protein with indole3-acetonitrile as substrate). This level of activity was lower than that found in A. thaliana, and this was deemed essential for the in vivo analysis. Leaf tissue from the transgenic plants converted 1-[13C]-indole-3-acetonitrile to 1-[13C]-indole-3-acetic acid in vivo as determined by HPLC/ GC-MS analysis. Untransformed tobacco was unable to catalyze this reaction. When transgenic seeds were grown on medium in the absence of indole-3-acetonitrile, germination and seedling growth appeared normal. In the presence of micromolar levels of exogenous indole-3-acetonitrile, a strong auxin-overproducing phenotype developed resulting in increased lateral root formation (at 10 µM indole-3-acetonitrile) or stunted shoot growth, excessive lateral root initiation, inhibition of root out-growth and callus formation at the root/shoot interface (at 100 µM indole-3-acetonitrile). Collectively, these data prove the ability of nitrilase II to convert low micromolar levels of indole-3-acetonitrile to indole-3-acetic acid in vivo, even when expressed at subphysiological levels thereby conferring a high-auxin phenotype upon transgenic plants. Thus, the A. thaliana nitrilase activity, which exceeds that of the transgenic plants, would be sufficient to meet the requirements for auxin biosynthesis in vivo.  相似文献   

10.
Abstract

Zofenopril as an ACE inhibitor expired recently was found to have a favourable safety profile in comparison with other ACE inhibitors in treating high blood pressure, congestive heart failure, and acute myocardial infarction. It can be synthesised from the key building blocks of (S)-3-benzoylthio-2-methylpropanoic acid and (4S)-phenylthio-L-proline. In this report, an efficient hydrolytic resolution via Candida antarctic lipase B (CALB) for preparing the former block in isopropyl ether (IPE) containing (RS)-3-benzoylthio-2-methylpropyl pyrazolide (1) and water was developed. Quantitative improvements of the enzyme activity and enantioselectivity in terms of k2SKmS?1?=?5.726?L h?1 g?1 and E?=?217 at 45?°C were found from the kinetic analysis. Insights into the CALB performance via thermodynamic analysis were then addressed and compared with those by using (RS)-3-benzoylthio-2-methylpropyl 1,2,4-triazolide (2) as the substrate. A putative thermodynamic model was moreover hypothesised for elucidating the more enthalpy reduction of 68.92-70.86?kJ mol?1 from the acyl part of (S)-1 and (S)-2 as well as that of 23.69-25.63?kJ mol?1 from the triad imidazolium to Ser105 and leaving 1,2,4-triazole moiety of (R)-2 and (S)-2 on stabilising the corresponding transition states.  相似文献   

11.
Summary

A supernatant from eggs of the ruminant nematode Trichostrongylus colubriformis contained an enzyme that was similar to leucine aminopeptidase (LAP), based on hydrolysis of the substrate L-leucine β-naphthylamide to β-naphthylamine. A Michaelis-Menten constant (K m) of 0.155 mM was obtained. Rate of hydrolysis of 16 substrates revealed that L-phenylalanine and L-tyrosine β-naphthylamides were hydrolyzed most readily while seven additional substrates were hydrolyzed at lesser rates. The optimum pH for enzymatic activity was 6.75–7.5. Enzymatic activity was lost by heating the egg supernatant to 60°C for 5 min or freezing at 0°C for 28 days. Addition of millimolar concentrations of the chlorides of zinc, manganese and magnesium to the egg supernatant had no stimulatory effect on enzyme activity while 10 and 100 mM concentrations significantly reduced activity. Ethylenediamine tetraacetic acid at 10?4 M had no effect on enzymatic activity. Activity was inhibited by 10?4 M 1,10-phenanthroline, but the inhibition was reversed by zinc chloride at 10?3 M. Di-isopropylphosphofluoridate at 10?3 M reduced enzymatic activity moderately. Enzyme activity in egg supernatant increased 2.2-fold from 21 days to 60–90 days of a primary infection in the host while a 3.3-fold increase was found in primary versus secondary infections.  相似文献   

12.
An aspartic protease that is significantly produced by baculovirus-infected Spodoptera frugiperda Sf9 insect cells was purified to homogeneity from a growth medium. To monitor aspartic protease activity, an internally quenched fluoresce (IQF) substrate specific to cathepsin D was used. The purified aspartic protease showed a single protein band on SDS–PAGE with an apparent molecular mass of 40 kDa. The N-terminal amino acid sequence of the enzyme had a high homology to a Bombyx mori aspartic protease. The enzyme showed greatest affinity for the IQF substrate at pH 3.0 with a K m of 0.85 μM. The k cat and k cat?K m values were 13 s?1 and 15 s?1 μM?1 respectively. Pepstatin A proved to be a potent competitive inhibitor with inhibitor constant, K i, of 25 pM.  相似文献   

13.
14.
External carbonic anhydrase (CA) activity in Chlorella saccharophila is suppressed by growth at high dissolved inorganic carbon and at acid pH. External CA activity was shown to be suppressed by growth at pHs below 7.0, with total repression at pH5.0. Growth in the presence of the buffer 3-[N-Morpholino]propane-sulphonic acid (MOPS) between pH 7 and 8 suppressed CA activity. Cells grown at pH8.0 aerated at 6 dm3 h?1 exhibited external CA activity of 5 units mg?1 Chl once the dissolved inorganic carbon (DIC) was reduced to 300 mmol m?3, and this increased to 30 units mg?1 Chl over a period of 3d while the DIC dropped to 30mmol m?3. Cells aerated at 180 dm3 h?1 showed a similar trend in CA activity, although the onset was delayed by 1 d and the DIC did not drop below 300 mmol m?3. Cells grown at pH 7.8 near an air equilibrium DIC of 300 mmol m?3had no detectable external CA activity. It is probable that it is the CO2 supply to the cell, and not total DIC or HCO?3 which controls external CA activity. Cells grown at pH 5.0 had no detectable activity, although they reduced the CO2 concentration to 0.6 mmol m?3. The loss of CA upon transfer of air-grown cells to 10 mmol mol?1 CO2 took place over 48 h and was light dependent, while the loss upon transfer from alkaline pH to acid pH look place over 12 h and was independent of light. The effects of pH are independent of the response to CO2.  相似文献   

15.
《Biomarkers》2013,18(4):267-272
Abstract

Sulphonamide hypersensitivity reactions are believed to be mediated through reactive intermediates derived from oxidation of the paraamino group to form sulphonamide hydroxylamines. Sulphamethoxazole hydroxylamine (SMX-HA) can be acetylated by N-acetyltransferase (NAT) enzymes to form an acetoxy metabolite (acetoxySMX). In the current studies, acetoxySMX was found to be not toxic over the concentration range of 0 to 500 μM towards a human lymphoblastoid cell line (RPMI 1788) or a human hepatoma cell line (HepG2). Further, transient expression of NAT1 in COS-1 cells or stable transfection of NAT1 andNAT2 in HepG2 cells did not alter the toxicity of SMX-HA in vitro. The activity of NAT1 in isolated mononuclear leucocytes (a reflection of systemic NAT1 activity) determined with paraaminobenzoic acid as a substrate was not different between controls (n = 11) or patients with a known hypersensitivity reaction (n = 5) (4.1 ±1.2 nmol min?1mg?1 vs 5.7 ± 1.4 nmol min?1 mg?1). Thus, acetoxy SMX is unlikely to play a significant role in mediating SMX hypersensitivity reactions anda constitutive deficiency in NAT1 activity is not a common finding in patients susceptible to SMX hypersensitivity reactions.  相似文献   

16.
Gas-vesicle (Vac) synthesis in Halobacterium salinarium PHH1 involves the expression of the p-vac region consisting of 14 different gvp genes that are arranged in two clusters: p-gvpACNO and, oppositely oriented, p-gvpDEFGHIJKLM. The latter cluster of genes is transcribed as two units: p-gvpDE and p-gvpF–M. The 5′-terminus of the p-gvpF–M mRMA was located 169 nucleotides upstream of p-gvpF within p-gvpE. The p-gvpG and p-gvpK gene was expressed in Escherichia coli and antibodies to proteins obtained were raised in rabbits. Both proteins could be detected in halobacterial cell lysates; in gas-vesicle preparations, however, neither GvpG nor GvpK could be found. The requirement for single p-gvp gene expression for gas-vesicle synthesis was determined by transformation experiments using the Vac? species Haloferax volcanii as recipient. Construct ΔA containing all p-gvp genes except for p-gvpA, encoding the major gas-vesicle structural protein, produced Vac? transformants, but the addition of p-gvpA on a second vector restored gas-vesicle synthesis to wild-type level (Vac++). Similarly, double transformants containing p-gvpD–M plus p-gvpACNO, or p-gvpG–M (fused to the promoter of the halobacterial ferredoxin gene for expression) plus p-gvpFED–ACNO were Vac++. Transformants containing the p-vac region either lacking gvpA, gvpF, or gvpGHI were Vac?, indicating the absolute requirement of these gvp genes (or at least one in the case of gvpGHI) for gas-vesicle formation. Double transformants containing the constructs p-gvpF–M plus p-gvpACNO (ΔDE) accumulated gas vesicles (Vac+) but synthesized fewer than the wild type, showing that the p-gvpDE genes are not necessary for gas-vesicle assembly. A repressor function affecting the synthesis of the p-gvpF–M mRNA could be suggested for p-gvpD and the 5′- region of its mRNA.  相似文献   

17.
Abstract

Oenococcus oeni CECT4730, which catalyses the asymmetric reduction of 2-octanone to (R)-2-octanol with high enantioselectivity, was further studied to exploit its potential for production of (R)-2-octanol in an aqueous/organic solvent biphasic system. Variables such as the volume ratio of aqueous to organic phase (Va/Vo), buffer pH, reaction temperature, shaking speed, co-substrates and the ratio of biocatalyst to substrate were examined with respect to the molar conversion, the initial reaction rate and the product enantiomeric excess (e.e.). Under the optimized conditions (Va/Vo=1:1 (v/v), buffer pH=8.0, reaction temperature=30°C, shaking speed=150 rev/min, ratio of glucose to biomass=5.4:l (w/w), ratio of biocatalyst to substrate=0.51:l (g/mol)), the highest space time yield of (R)-2-octanol, 24 mmol L?1 per h, and >98% product e.e. were obtained at a substrate concentration close to 1.0 mol L?1 after 24 h reduction.  相似文献   

18.
Microbial nitrilases are biocatalysts of interest and the enzyme produced using various inducers exhibits altered substrate specificity, which is of great interest in bioprocess development. The aim of the present study is to investigate the nitrilase-producing Alcaligenes faecalis MTCC 10757 (IICT-A3) for its ability to transform various nitriles in the presence of different inducers after optimization of various parameters for maximum enzyme production and activity. The production of A. faecalis MTCC 10757 (IICT-A3) nitrilase was optimum with glucose (1.0%), acrylonitrile (0.1%) at pH 7.0. The nitrilase activity of A. faecalis MTCC 10757 (IICT-A3) was optimum at 35 °C, pH 8.0 and the enzyme was stable up to 6 h at 50 °C. The nitrilase enzyme produced using different inducers was investigated for substrate specificity. The enzyme hydrolyzed aliphatic, heterocyclic and aromatic nitriles with different substitutions. Acrylonitrile was the most preferred substrate (~40 U) as well as inducer. Benzonitrile was hydrolyzed with almost twofold higher relative activity than acrylonitrile when it was used as an inducer. The versatile nitrilase-producing A. faecalis MTCC 10757 (IICT-A3) exhibits efficient conversion of both aliphatic and aromatic nitriles. The aromatic nitriles, which show not much or no affinity towards nitrilase from A. faecalis, are hydrolyzed effectively with this nitrilase-producing organism. Studies are in progress to exploit this organism for synthesis of industrially important compounds.  相似文献   

19.
Biotransformation of the antifungal compound 16-oxocleroda-3,13(14)E-dien-15-oic acid (1) isolated from Polyalthia longifolia leaves was achieved by Rhizopus stolonifer in broth medium containing the substrate at the sublethal concentration of 0.06?mg ml?1. A novel derivative, 18-hydroxy-16-oxocleroda-3,13(14)E-dien-15-oic acid (2) was isolated after 4 d of incubation. Minimum inhibitory concentration (MIC) was determined against 11 fungal pathogens of clinical and agricultural importance. The biotransformed compound showed lower MIC values than the natural parent compound. The study showed that the fungus R. stolonifer has the potential to hydroxylate a natural fungicidal clerodane diterpene at allylic position to produce a novel hydroxylated derivative with enhanced antifungal activity.  相似文献   

20.
A series of substituted 2,4,5-triphenylisothiazol-3(2H)-one 1,1-dioxides 9 was synthesized and investigated as inhibitors of human leukocyte elastase (HLE). All compounds were found to inhibit HLE in a time-dependent manner and most of them exhibited kobs/[I] values > 300 M? 1s? 1. The most potent 3-oxosultam of this series was 9l (kobs/[I] = 2440 M? 1s? 1). Kinetic investigations performed with 9g and different substrate concentrations did not allow to clearly distinguish between a competitive or noncompetitive mode of inhibition. A more complex interaction is supported by the failure of a linear dependency of kobs values on the inhibitor concentration.  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