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.
Environmental contamination by hexavalent chromium, Cr(VI), presents a serious public health problem. This study assessed the reduction of Cr(VI) by intact cells and a cell-free extract (CFE) of an actinomycete, Arthrobacter crystallopoietes (strain ES 32), isolated from soil contaminated with dichromate. Both intact cells and CFE of A. crystallopoietes, displayed substantial reduction of Cr(VI). Intact cells reduced about 90% of the Cr(VI) added within 12 h and Cr(VI) was almost completely reduced after 24 h. The K M and V max of Cr(VI) bioreduction by intact cells were 2.61 μM and 0.0142 μmol/min/mg protein, respectively. Cell-free chromate reductase of the A. crystallopoietes (ES 32) reduced hexavalent chromium at a K M of 1.78 μM and a V max of 0.096 μmol/min/mg protein. The rate constant (k) of chromate reduction was inversely related to Cr(VI) concentration and the half-life (t 1/2) of Cr(VI) reduction increased with increasing concentration. A. crystallopoietes produced a periplasmic chromate reductase that was stimulated by NADH. Results indicate that A. crystallopoietes ES 32 can be used to detoxify Cr(VI) in polluted sites, particularly in stressed environments.  相似文献   

2.
Bacillus sp. ES 29 (ATCC: BAA-696) is an efficient chromate reducing bacterium. We evaluated hexavalent chromium (Cr[VI]) reduction by immobilized intact cells and the cell-free enzyme extracts of Bacillus sp. ES 29 in a bioreactor system. Influences of different flow rates (3 to 14 mL h?1), Cr(VI) concentration (2 to 8 mg L?1), and immobilization support materials (Celite, amberlite, and Ca-alginate) on Cr(VI) reduction were examined. Both immobilized intact cells and the cell-free extract of Bacillus sp. ES 29 displayed substantial Cr(VI) reduction. Increasing flow rates from 3 to 6 mL h?1 did not affect the rate of Cr(VI) reduction, but above 6 mL h?1, the Cr(VI) reducing capacity of the immobilized intact cells and cell-free extract of Bacillus sp. ES 29 decreased. With both intact cells and the cell-free extracts, the rate of Cr(VI) reduction was inversely related to the concentration. Intact cells immobilized to Celite displayed the highest rate (k = 0.443 at 3 mL h?1) of Cr(VI) reduction. For the immobilized cell-free extract, maximal reduction (k = 0.689 at 3 mL h?1) was observed with Ca-alginate. Using initial Cr(VI) concentrations of 2 to 8 mg L?1 at flow rates of 3 to 6 mL h?1 both immobilized intact cells and the cell-free extracts reduced 84 to 98% of the influent Cr(VI). Results indicate that immobilized cells and the cell-free extracts of Bacillus sp. ES 29 could be used for large-scale removal of Cr(VI) from contaminated water and waste streams in containment systems.  相似文献   

3.
Bacillus sp. ES 29 (ATCC: BAA-696) is an efficient chromate reducing bacterium. We evaluated hexavalent chromium (Cr[VI]) reduction by immobilized intact cells and the cell-free enzyme extracts of Bacillus sp. ES 29 in a bioreactor system. Influences of different flow rates (3 to 14 mL h-1), Cr(VI) concentration (2 to 8 mg L-1), and immobilization support materials (Celite, amberlite, and Ca-alginate) on Cr(VI) reduction were examined. Both immobilized intact cells and the cell-free extract of Bacillus sp. ES 29 displayed substantial Cr(VI) reduction. Increasing flow rates from 3 to 6 mL h-1 did not affect the rate of Cr(VI) reduction, but above 6 mL h-1, the Cr(VI) reducing capacity of the immobilized intact cells and cell-free extract of Bacillus sp. ES 29 decreased. With both intact cells and the cell-free extracts, the rate of Cr(VI) reduction was inversely related to the concentration. Intact cells immobilized to Celite displayed the highest rate (k = 0.443 at 3 mL h-1) of Cr(VI) reduction. For the immobilized cell-free extract, maximal reduction (k = 0.689 at 3 mL h-1) was observed with Ca-alginate. Using initial Cr(VI) concentrations of 2 to 8 mg L-1 at flow rates of 3 to 6 mL h-1 both immobilized intact cells and the cell-free extracts reduced 84 to 98% of the influent Cr(VI). Results indicate that immobilized cells and the cell-free extracts of Bacillus sp. ES 29 could be used for large-scale removal of Cr(VI) from contaminated water and waste streams in containment systems.  相似文献   

4.
A pseudomonad (CRB5) isolated from a decommissioned wood preservation site reduced toxic chromate [Cr(VI)] to an insoluble Cr(III) precipitate under aerobic and anaerobic conditions. CRB5 tolerated up to 520 mg of Cr(VI) liter(-1) and reduced chromate in the presence of copper and arsenate. Under anaerobic conditions it also reduced Co(III) and U(VI), partially internalizing each metal. Metal precipitates were also found on the surface of the outer membrane and (sometimes) on a capsule. The results showed that chromate reduction by CRB5 was mediated by a soluble enzyme that was largely contained in the cytoplasm but also found outside of the cells. The crude reductase activity in the soluble fraction showed a K(m) of 23 mg liter(-1) (437 microM) and a V(max) of 0.98 mg of Cr h(-1) mg of protein(-1) (317 nmol min(-1) mg of protein(-1)). Minor membrane-associated Cr(VI) reduction under anaerobiosis may account for anaerobic reduction of chromate under nongrowth conditions with an organic electron donor present. Chromate reduction under both aerobic and anaerobic conditions may be a detoxification strategy for the bacterium which could be exploited to bioremediate chromate-contaminated or other toxic heavy metal-contaminated environments.  相似文献   

5.
A soluble Cr(VI) reductase was purified from the cytoplasm of Escherichia coli ATCC 33456. The molecular mass was estimated to be 84 and 42 kDa by gel filtration and SDS-polyacrylamide gel electrophoresis, respectively, indicating a dimeric structure. The pI was 4.66, and optimal enzyme activity was obtained at pH 6.5 and 37 degrees C. The most stable condition existed at pH 7.0. The purified enzyme used both NADPH and NADH as electron donors for Cr(VI) reduction, while NADPH was the better, conferring 61%; higher activity than NADH. The Km values for NADPH and NADH were determined to be 47.5 and 17.2 micromol, and the Vmax values 322.2 and 130.7 micromol Cr(VI) min(-1)mg(-1) protein, respectively. The activity was strongly inhibited by N-ethylmalemide, Ag2+, Cd2+, Hg2+, and Zn2+. The antibody against the enzyme showed no immunological cross reaction with those of other Cr(VI) reducing strains.  相似文献   

6.
Three efficient Cr(VI) reducing bacterial strains were isolated from Cr(VI) polluted landfill and characterized for in vitro Cr(VI) reduction. Phylogenetic analysis using 16S rRNA gene sequencing revealed that the newly isolated strains G1DM20, G1DM22 and G1DM64 were closely related to Bacillus cereus, Bacillus fusiformis and Bacillus sphaericus, respectively. The suspended cultures of all Bacillus sp. exhibited more than 85% reduction of 1000 microM Cr(VI) within 30 h. The suspended culture of Bacillus sp. G1DM22 exhibited an ability for continuous reduction of 100 microM Cr(VI) up to seven consecutive inputs. Assays with the permeabilized cells and cell-free extracts from each of Bacillus sp. demonstrated that the hexavalent chromate reductase activity was mainly associated with the soluble fraction of cells and expressed constitutively. The Cr(VI) reduction by the cell-free extracts of Bacillus sp. G1DM20 and G1DM22 was maximum at 30 degrees C and pH 7 whereas, Bacillus sp. G1DM64 exhibited maximum Cr(VI) reduction at pH 6. Addition of 1mM NADH enhanced the Cr(VI) reductase activity in the cell-free extracts of all three isolates. Amongst all three isolates tested, crude cell-free extracts of Bacillus sp. G1DM22 exhibited the fastest Cr(VI) reduction rate with complete reduction of 100 microM Cr(VI) within 100 min. The apparent K(m) and V(max) of the chromate reductase activity in Bacillus sp. G1DM22 were determined to be 200 microM Cr(VI) and 5.5 micromol/min/mg protein, respectively. The Cr(VI) reductase activity in cell-free extracts of all the isolates was stable in presence of different metal ions tested except Hg(2+) and Ag(+).  相似文献   

7.
The reduction of hexavalent chromium (Cr(VI] by rat liver microsomes was studied. With 15-120 microM Na2CrO4 microsomes (0.5 mg protein/ml) effectively reduced Cr(VI) in the presence of NADPH provided anaerobic conditions. Phenobarbital (PB) and Aroclor 1254 (PCB) pretreatment increased microsomal Cr(VI) reduction while CoCl2 reduced the rate. The rates with 30 microM Na2CrO4 were: 6.4 +/- 0.1, 7.8 +/- 0.7, 13.4 +/- 0.5, 2.95 +/- 0.09 nmol Cr.mg prot.-1 min-1 for control, PB, PCB and cobalt pretreated microsomes respectively. Kinetic studies gave a Michaeli-Menten like first-order kinetics with increases both in Km and Vmax values after pretreatment with PB or PCB. CO partly inhibited the microsomal Cr(VI) reduction. The CO-sensitive reduction rate was directly correlated to the cyt. P-450 content of the different microsomal preparations. Substituting NADH for NADPH gave approximately 27% lower activity with 30 microM Na2CrO4. This activity was neither inducible by cyt. P-450 inducers nor influenced by CO. Oxygen 1.0% and 0.10% gave approximately 100% and 30% inhibition of Cr(VI) reduction (30 microM Na2CrO4) respectively, and an uncompetitive like inhibitory pattern was found. No redox cycling of Cr(VI) was seen. 51Cr binding to the microsomes was approximately 10% after complete reduction of 30 microM Na2CrO4. Externally added FMN, Fe3+-ADP and nitrobenzen stimulated microsomal Cr(VI) reduction. A 60% higher reduction rate of Cr(VI) by isolated hepatocytes was found during anaerobic in comparison with aerobic conditions.  相似文献   

8.
Microbial enzymatic reduction of a toxic form of chromium [Cr(VI)] has been considered as an effective method for bioremediation of this metal. This study reports on the in vitro reduction of Cr(VI) using cell-free extracts from a Cr(VI) reducing Bacillus firmus KUCr1 strain. Chromium reductase was found to be constitutive and its activity was observed both in soluble cell fractions (S12 and S150 and membrane cell fraction (P150). The reductase activity of S12 fraction was found to be optimal at 40 microM Cr(VI) with enzyme concentration equivalent to 0.493 mg protein/ml. Enzyme activity was dependent on NADH or NADPH as electron donor; optimal temperature and pH for better enzyme activity were 70 degrees C and 5.6, respectively. The Km value of the reductase was 58.33 microM chromate having a V(max) of 11.42 microM/min/mg protein. The metabolic inhibitor like sodium azide inhibited reductase activity of membrane fraction of the cell-free extract. Metal ions like Cu2+, Co2+, Ni2+ and As3+ stimulated the enzyme but others, such as Ag+, Hg2+, Zn2+, Mn2+, Cd2+ and Pb2+, inhibited Cr(VI) reductase activity.  相似文献   

9.
The apicomplexan Cryptosporidium parvum possesses a unique 1500-kDa polyketide synthase (CpPKS1) comprised of 29 enzymes for synthesising a yet undetermined polyketide. This study focuses on the biochemical characterization of the 845-amino acid loading unit containing acyl-[ACP] ligase (AL) and acyl carrier protein (ACP). The CpPKS1-AL domain has a substrate preference for long chain fatty acids, particularly for the C20:0 arachidic acid. When using [3H]palmitic acid and CoA as co-substrates, the AL domain displayed allosteric kinetics towards palmitic acid (Hill coefficient, h=1.46, K50=0.751 microM, Vmax=2.236 micromol mg(-1) min(-1)) and CoA (h=0.704, K50=5.627 microM, Vmax=0.557 micromol mg(-1) min(-1)), and biphasic kinetics towards adenosine 5'-triphosphate (Km1=3.149 microM, Vmax1=373.3 nmol mg(-1) min(-1), Km2=121.0 microM, and Vmax2=563.7 nmol mg(-1) min(-1)). The AL domain is Mg2+-dependent and its activity could be inhibited by triacsin C (IC50=6.64 microM). Furthermore, the ACP domain within the loading unit could be activated by the C. parvum surfactin production element-type phosphopantetheinyl transferase. After attachment of the fatty acid substrate to the AL domain for conversion into the fatty-acyl intermediate, the AL domain is able to transfer palmitic acid to the activated holo-ACP in vitro. These observations ultimately validate the function of the CpPKS1-AL-ACP unit, and make it possible to further dissect the function of this megasynthase using recombinant proteins in a stepwise procedure.  相似文献   

10.
Cell-free extracts (CFEs) of chromium-resistant bacterium Bacillus sphaericus AND 303 isolated from serpentine soil of Andaman, India reduced Cr(VI) in in vitro condition, and the reductase activity was solely localized in the soluble cell-fractions (S12, S32, and S150). The enzyme was constitutive as the CFEs from cells grown in Cr(VI)-free and Cr(VI)-containing media reduced a more or less equal amount of Cr(VI). Optimum Cr(VI) reductase activity was obtained at an enzyme (S150) concentration equivalent to 4.56 mg protein/mL, 300 μM Cr(VI) and pH 6.0 after 30 min incubation at 30°C. The enzyme was heat labile; 80% of its activity was lost when exposed at 70°C for 15 min. Kinetics of Cr(VI) reductase activity fit well with the linearized Lineweaver-Burk plot and showed a Vmax of 1.432 μmol Cr(VI)/mg protein/min and Km of 158.12 μM Cr(VI). The presence of additional electron donors accelerated Cr(VI) reductase activity of CFE, and an increase of 28% activity over control was recorded with 1.0 μM NADH. Heavy metal ions such as Ni(II), Cu(II), and Cd(II) were strong inhibitors of Cr(VI) reductase unlike that of 100 μM Co(II), which retained 93% activity over control.  相似文献   

11.
A cyanide-degrading enzyme from Bacillus pumilus C1 has been purified and characterized. This enzyme consisted of three polypeptides of 45.6, 44.6, and 41.2 kDa; the molecular mass by gel filtration was 417 kDa. Electron microscopy revealed a multimeric, rod-shaped protein approximately 9 by 50 nm. Cyanide was rapidly degraded to formate and ammonia. Enzyme activity was optimal at 37 degrees C and pH 7.8 to 8.0. Activity was enhanced by Sc3+, Cr3+, Fe3+, and Tb3+; enhancement was independent of metal ion concentration at concentrations above 5 microM. Reversible enhancement of enzymatic activity by azide was maximal at 4.5 mM azide and increased with time. No activity was recorded with the cyanide substrate analogs CNO-, SCN-, CH3CN, and N3- and the possible degradation intermediate HCONH2. Kinetic studies indicated a Km of 2.56 +/- 0.48 mM for cyanide and a Vmax of 88.03 +/- 4.67 mmol of cyanide per min/mg/liter. The Km increased approximately twofold in the presence of 10 microM Cr3+ to 5.28 +/- 0.38 mM for cyanide, and the Vmax increased to 197.11 +/- 8.51 mmol of cyanide per min/mg/liter. We propose naming this enzyme cyanide dihydratase.  相似文献   

12.
Succinate:quinone oxidoreductase (EC 1.3.5.1) was first purified from the facultative alkaliphilic Bacillus sp. strain YN-2000 in the presence of Triton X-100. The isolated enzyme showed high succinate-ubiquinone oxidoreductase activity at pH 8.5. The Km for ubiquinone 1 and the Vmax of the enzyme were determined to be about 5 microM and 48 micromol of ubiquinone 1 per min per mg, respectively. The catalytic activity of the enzyme was 50% inhibited by 9 microM 2-thenoyltrifluoroacetone or 0.8 microM 2-n-heptyl-4-hydroxyquinoline- N-oxide. The enzyme consisted of three kinds of subunits with molecular masses of 66, 26, and 15 kDa, respectively, and contained 1.28 mol of covalently bound flavin adenine dinucleotide, 0.9 mol of heme b, 1.35 mol of menaquinone, 8.3 mol of nonheme iron, and 7.5 mol of inorganic sulfide per mol of enzyme. The enzyme showed symmetrical alpha absorption peaks at 556.5 and 554 nm in the reduced state at room temperature and 77 K, respectively. The potentiometric analysis of the enzyme yielded an Em,7 of heme b of about -64 mV (n = 1). Furthermore, the content of the enzyme was increased up to fivefold when the bacterium was grown at pH 10 compared with pH 7. These results indicate that the succinate:quinone oxidoreductase with a single heme b is involved in the respiratory chain of the alkaliphile at a very alkaline pH.  相似文献   

13.
Folylpolyglutamate synthetase (FPGS) was isolated from human liver cytosol by 0-30% (w/v) ammonium sulfate fractionation and characterized biochemically. Using aminopterin (AMT), L-[3H]glutamate and MgATP as cosubstrates, maximal gamma-L-glutamylation activity was observed in the presence of the activators KCl and NaHCO3. ATP and 2-mercaptoethanol were each required for enzyme activity and stability. In the absence of ATP, human liver FPGS rapidly inactivated at 37 degrees C (t1/2 approximately 8 min), whereas FPGS isolated from rabbit liver was significantly more stable (t1/2 = 68 min). Both folates and antifolates were effectively polyglutamylated by the isolated human liver enzyme. Km parameters determined for AMT (Km = 4.3 microM) were similar to those determined for several reduced folates (tetrahydrofolic acid, dihydrofolic acid, and folinic acid; Km = 3-7 microM), while significantly higher Km values were observed for methotrexate (MTX) and 5-methyltetrahydrofolic acid (Km = 50-60 microM) and for folic acid (Km = 100 microM). All of the substrates examined exhibited Vmax values ranging from 30 to 90% of the AMT value (Vmax = 935 pmol product/mg/h). The order of reactivity for these substrates differed from that determined in parallel studies for FPGS isolated from rat and rabbit liver. In the case of AMT and several reduced folates, inhibition of human liver FPGS was observed at substrate concentrations at or above 50-250 microM. FPGS isolated from six individual human livers exhibited highly similar biochemical and kinetic properties, suggesting the presence of the same or at least highly similar enzyme species in each individual, with a five-fold interindividual range in specific activities observed. Comparison of MTX with its higher polyglutamates (MTX-Glu2 to MTX-Glu6) as FPGS substrates indicated a significant decrease in Vmax values with increasing glutamate chain length which was partially compensated for by a corresponding decrease in Km. Consistent with these observations, the isolated enzyme was unable to synthesize polyglutamates higher than MTX-Glu3 when MTX was supplied as substrate, raising the question as to how MTX polyglutamates containing up to five or six gamma-L-glutamate residues are formed in vivo.  相似文献   

14.
We analyzed the kinetics and metabolic pathways of trichloroethylene and 1,1,1-trichloroethane degradation by the ethane-utilizing Mycobacterium sp. TA27. The apparent Vmax and Km of trichloroethylene were 9.8 nmol min(-1) mg of cells(-1) and 61.9 microM, respectively. The apparent Vmax and Km of 1,1,1-trichloroethane were 0.11 nmol min(-1) mg of cells(-1) and 3.1 microM, respectively. 2,2,2-trichloroethanol, trichloroacetic acid, chloral, and dichloroacetic acid were detected as metabolites of trichloroethylene. 2,2,2-trichloroethanol, trichloroacetic acid, and dichloroacetic acid were also detected as metabolites of 1,1,1-trichloroethane. The amounts of 2,2,2-trichloroethanol, trichloroacetic acid, chloral, and dichloroacetic acid derived from the degradation of 3.60 micromol trichloroethylene were 0.16 micromol (4.4%), 0.11 micromol (3.1%), 0.02 micromol (0.6%), and 0.02 micromol (0.6%), respectively. The amounts of 2,2,2-trichloroethanol, trichloroacetic acid and dichloroacetic acid derived from the degradation of 1.73 micromol 1,1,1-trichloroethane were 1.48 micromol (85.5%), 0.22 micromol (12.7%), and 0.02 micromol (1.2%), respectively. More than 90% of theoretical total chloride was released in trichloroethylene degradation. Chloral and 2,2,2-trichloroethanol were transformed into each other, and were finally converted to trichloroacetic acid, and dichloroacetic acid. Trichloroacetic acid and dichloroacetic acid were not degraded by strain TA27.  相似文献   

15.
Two species of Pseudomonas capable of utilizing nitroglycerin (NG) as a sole nitrogen source were isolated from NG-contaminated soil and identified as Pseudomonas putida II-B and P. fluorescens I-C. While 9 of 13 laboratory bacterial strains that presumably had no previous exposure to NG could degrade low concentrations of NG (0.44 mM), the natural isolates tolerated concentrations of NG that were toxic to the lab strains (1.76 mM and higher). Whole-cell studies revealed that the two natural isolates produced different mixtures of the isomers of dinitroglycerol (DNG) and mononitroglycerol (MNG). A monomeric, flavin mononucleotide-containing NG reductase was purified from each natural isolate. These enzymes catalyzed the NADPH-dependent denitration of NG, yielding nitrite. Apparent kinetic constants were determined for both reductases. The P. putida enzyme had a Km for NG of 52 +/- 4 microM, a Km for NADPH of 28 +/- 2 microM, and a Vmax of 124 +/- 6 microM x min(-1), while the P. fluorescens enzyme had a Km for NG of 110 +/- 10 microM, a Km for NADPH of 5 +/- 1 microM, and a Vmax of 110 +/- 11 microM x min(-1). Anaerobic titration experiments confirmed the stoichiometry of NADPH consumption, changes in flavin oxidation state, and multiple steps of nitrite removal from NG. The products formed during time-dependent denitration reactions were consistent with a single enzyme being responsible for the in vivo product distributions. Simulation of the product formation kinetics by numerical integration showed that the P. putida enzyme produced an approximately 2-fold molar excess of 1,2-DNG relative to 1,3-DNG. This result could be fortuitous or could possibly be consistent with a random removal of the first nitro group from either the terminal (C-1 and C-3) positions or middle (C-2) position. However, during the denitration of 1,2-DNG, a 1.3-fold selectivity for the C-1 nitro group was determined. Comparable simulations of the product distributions from the P. fluorescens enzyme showed that NG was denitrated with a 4.6-fold selectivity for the C-2 position. Furthermore, a 2.4-fold selectivity for removal of the nitro group from the C-2 position of 1,2-DNG was also determined. The MNG isomers were not effectively denitrated by either purified enzyme, which suggests a reason why NG could not be used as a sole carbon source by the isolated organisms.  相似文献   

16.
Bacillus subtilis exhibited an inducible K+-transporting ATPase activity with apparent Km and maximum velocity Vmax of 12.9 microM and 25.1 micromol x min(-1) x (g cell protein)(-1), respectively, when cultivated on a synthetic medium containing less than 400 microM K+. Due to this enzyme, the growth rate of the bacterium in synthetic medium was not changed down to 115 microM K+, and the bacterium was able to grow down to 20 microM K+. The limiting K+ concentration was higher in media with osmolarity increased by NaCl or sucrose. The ATPase was inhibited by micromolar concentrations of vanadate (Ki = 1.6 microM). The ATPase activity was not stimulated by any other monovalent cation. The subunit of this ATPase, with an Mr of 52000, covalently bound the gamma phosphate group of ATP. This phosphorylated intermediate was unstable in neutral and basic pH as well as in the presence of potassium and was stable in acid pH. The enzyme did not show immunological cross-reactivity with antibody against Kdp ATPase of Escherichia coli.  相似文献   

17.
Hexavalent chromium, Cr(VI), is toxic to living systems. Widespread contamination of water and soil by Cr(VI) present a serious public health problem. Chromium-resistant bacteria can reduce and detoxify Cr(VI). Twelve bacteria resistant to high concentrations of Cr(VI) were isolated from soil enrichment cultures. Environmental parameters and kinetic parameters of Cr(VI) bioreduction by one monoculture isolate, identified by 16S rRNA gene sequence as Bacillus sp. PB2, were studied. The optimal temperature for growth and Cr(VI) reduction was 35 degrees C. The isolate grew luxuriantly and substantially reduced Cr(VI) at initial pH 7.5 to 9. Maximal Cr(VI) bioreduction occurred at initial pH 8.0. Substantial Cr(VI) bioreduction was observed in salt media, but removal efficiency was inversely related to salt concentration (1-9%). Michaelis-Menten hyperbolic equation and the Lineweaver-Burk double reciprocal plot were comparatively employed to determine the k (m) and V (max) of Cr(VI) bioreduction. A k (m) of 82.5 microg mL(-1) and V (max) of 7.78 microg mL(-1) h(-1) were calculated by nonlinear regression analysis of the hyperbola curve. Linear regression analysis of the double reciprocal plot revealed k (m) and V (max) of 80.9 microg mL(-1) and 10.6 microg mL(-1) h(-1), respectively. Time course studies displayed about 90% reduction of Cr(VI) at an initial concentration of 8,000 microg L(-1) in 8 h, with an estimated t (1/2) of 4 h. Data from time course analysis of the rate of Cr(VI) bioreduction fitted zero-order model, and the kinetic constant k was calculated to be 840 microg L(-1) h(-1). The monoculture isolate, Bacillus sp. PB2, strongly reduces Cr(VI) and could be used for bioremediation of Cr(VI)-contaminated aquatic and terrestrial environments.  相似文献   

18.
Chromate resistant and reducing strains were isolated from chromium contaminated soil and identified as Bacillus sp. (KCH2 and KCH3), Leucobacter sp. (KCH4) and Exiguobacterium sp. (KCH5). KCH3 and KCH4 showed higher Cr(VI) tolerance (2 mM) and Cr(VI) reduction (1.5 mM) than KCH5 (1.5 mM and 0.75 mM, respectively). Cr(VI) reduction by CFEs of KCH3 and KCH4 showed NAD(P)H dependence, optimum activity at pH 5.5, low K(m) (45-55 microM) and substrate inhibition by Cr(VI) (>75 microM), whereas that of KCH5 showed NADH dependence, pH optimum at 6.0, high K(m) (200 microM) and no inhibition by Cr(VI). Cr(VI) reduction was optimum at 35 degrees C for CFEs of KCH3 and KCH5 and 30 degrees C for that of KCH3. Cr(VI) reduction by CFEs of all the strains were inhibited by Hg(2+) and enhanced by Cu(2+). Activity enhancement by Cu(2+) was more predominant (290%) for KCH4. The characterization of Cr(VI) reduction by CFEs of chromate resistant isolates of different genera is useful for development of Cr(VI) bioremediation.  相似文献   

19.
In earlier studies, two distinct molecules, 20 alpha-HSD-I and 20 alpha-HSD-II, responsible for 20 alpha-HSD activity of pig adrenal cytosol were purified to homogeneity and characterized [S. Nakajin et al., J. Steroid Biochem. 33 (1989) 1181-1189]. We report here that the purified 20 alpha-HSD-I, which mainly catalyzes the reduction of 17 alpha-hydroxyprogesterone to 17 alpha,20 alpha-dihydroxy-4-pregnen-3-one, catalyzes 3 alpha-hydroxysteroid oxidoreductase activity for 5 alpha (or 5 beta)-androstanes (C19), 5 alpha (or 5 beta)-pregnanes (C21) in the presence of NADPH as the preferred cofactor. The purified enzyme has a preference for the 5 alpha (or 5 beta)-androstane substrates rather than 5 alpha (or 5 beta)-pregnane substrates, and the 5 beta-isomers rather than 5 alpha-isomers, respectively. Kinetic constants in the reduction for 5 alpha-androstanedione (Km; 3.3 microM, Vmax; 69.7 nmol/min/mg) and 5 beta-androstanedione (Km; 7.7 microM, Vmax; 135.7 nmol/min/mg) were demonstrated for comparison with those for 17 alpha-hydroxyprogesterone (Km; 26.2 microM, Vmax; 1.3 nmol/min/mg) which is a substrate for 20 alpha-HSD activity. Regarding oxidation, the apparent Km and Vmax values for 3 alpha-hydroxy-5 alpha-androstan-17-one were 1.7 microM and 43.2 nmol/min/mg, and 1.2 microM and 32.1 nmol/min/mg for 3 alpha-hydroxy-5 beta-androstan-17-one, respectively. 20 alpha-HSD activity in the reduction of 17 alpha-hydroxyprogesterone catalyzed by the purified enzyme was inhibited competitively by addition of 5 alpha-DHT with a Ki value of 2.0 microM. Furthermore, 17 alpha-hydroxyprogesterone inhibited competitively 3 alpha-HSD activity with a Ki value of 150 microM.  相似文献   

20.
In recent years, more and more attentions are put on the remediation of Cr(VI) contamination with chromate resistant bacteria. Leucobacter sp. CRB1 was a novel chromate reducing bacteria isolated from the soil of chromite ore processing residue (COPR) disposal site in Changsha, China. The objectives of this study were to evaluate the Cr(VI) tolerance of Leucobacter sp. CRB1 as well as its tolerant mechanism, and Cr(VI) reduction ability. The results showed that Leucobacter sp. CRB1 was able to tolerate 4,000 mg/l of hexavalent chromium with 34.5% reduction efficiency. At the optimum pH 9.0, the maximum concentration of chromate be reduced completely was 1,818 mg/l in growing cells and 2,100 mg/l in resting cells. Scanning electron microscopy (SEM) and energy dispersive X-ray analysis (EDAX) showed that extracellular Cr(VI) reduction of Leucobacter sp. CRB1 contributed to its high tolerance and high reduction ability. With repeating spiking, 2,490 mg/l hexavalent chromium was reduced totally within 17 h. The results suggest Leucobacter sp. CRB1 has potential application for remediation of high concentration of Cr(VI) contamination.  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